Next Article in Journal
Agro-Waste Derived Biomass Impregnated with TiO2 as a Potential Adsorbent for Removal of As(III) from Water
Previous Article in Journal
Heterogeneous Catalytic Ozonation of Phenol by a Novel Binary Catalyst of Fe-Ni/MAC
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Cu-Promoted Iron Catalysts Supported on Nanorod-Structured Mn-Ce Mixed Oxides for Higher Alcohol Synthesis from Syngas

1
Engineering Research Center of Large Scale Reactor Engineering and Technology, Ministry of Education, State Key Laboratory of Chemical Engineering, School of Chemical Engineering, East China University of Science and Technology, Shanghai 200237, China
2
State Key Laboratory of Coal Liquefaction and Coal Chemical Technology, Shanghai 201203, China
3
Department of Chemical Engineering, Norwegian University of Science and Technology, 7491 Trondheim, Norway
*
Author to whom correspondence should be addressed.
Catalysts 2020, 10(10), 1124; https://doi.org/10.3390/catal10101124
Submission received: 1 September 2020 / Revised: 27 September 2020 / Accepted: 28 September 2020 / Published: 1 October 2020

Abstract

:
A series of supports were prepared through the method of hydrothermal synthesis, and copper–iron catalysts supported on ceria nanorods modified by different amounts of manganese were prepared by the liquid phase co-reduction method. The effect of the catalytic performance after Mn addition mainly on higher alcohols synthesis (HAS) was evaluated. Different techniques, such as BET, ICP-AES, XRD, H2-TPR, CO-TPD, TEM, FESEM, XPS and MES, were performed for catalyst characterization. The results indicated that the abilities of CO chemical desorption and carbon chain growth were promoted with appropriate Mn addition, and higher ratio of Cu0/Cu+ species facilitated the methanol homologous reaction and the C2+OH formation. The Ce4+ species were reduced into Ce3+ species during HAS process, providing a large amount of oxygen vacancies. Proper Mn content promoted the formation of χ-Fe5C2 and leaded to the Fe 2p binding energy shift, causing the electron transformation between Fe and Mn species. The largest weight selectivity of C2+OH appeared in the reaction over CuFe/3.6MnCe catalyst with CO conversion 41.43%, and weight fraction of C2+OH 84.41 wt% in the alcohols distribution.

Graphical Abstract

1. Introduction

In recent years, with the decreasing reserve and rising price of petroleum resources, the consumption structure of energy and chemical industry has gradually changed from petroleum to the co-supply of petroleum, coal, natural gas and biomass. Fischer–Tropsch synthesis (FTS) is a reaction process with syngas (CO and H2) as raw materials to synthesize lots of products including hydrocarbons and high-valued alcohols under certain conditions [1,2]. Higher alcohols (C2+OH) can be used as transportation fuels to enhance octane number and improve engine performance, and intermediates to produce detergents and surfactants [3,4,5].
Many great efforts have been made to develop catalysts for FTS and higher alcohols synthesis, such as Fe, Co, Cu, MoS2 and Ru-based catalysts [1,6]. Compared with other catalysts, low cost and high chain growth possibility are the advantages of iron-based catalysts in FTS, but the selectivity of alcohols is much lower than that of modified Cu-based catalysts [7,8,9]. Due to the ability of facilitating carbon chain growth and achieving high selectivity of higher alcohols (with a selectivity of C2+OH larger than 20%), modified Cu-Fe catalysts have attracted more attentions in recent years [10,11,12]. The reaction over modified Cu-Fe catalysts is a bit different from the traditional FTS which focuses on the high selectivity of C5+ products or olefins and neglects the formation of alcohols, the synergistic effect of iron and copper, which combines the carbon chain growth effect of iron carbides and alcohols formation effect of copper together, could largely promote the formation of higher alcohols. On the other hand, as a basic chemical raw material, the economic value of higher alcohols is much higher than hydrocarbons, so the increase in the selectivity of higher alcohols can efficiently improve the economy of FTS process. Mn promoter added on catalysts can enhance the synergistic effect of iron and copper species in Fischer–Tropsch synthesis, which will facilitate the formation of Fe-Mn-O solid solution and achieving highly dispersed iron and copper species [12]. It was reported that Ce promoter could enhance the ability of CO chemical desorption and improve the carbonization because of its electronic donating effect, which will finally promote the formation of FeCx species and decrease the selectivity of methane (from 11.1% to 9.5% with Ce addition from 0 to 0.5%) [6]. The results of comparing the reaction performance of Cu-Fe, Cu-Co and Cu-Ni nanoparticles in FTS showed that Cu-Fe catalysts had best performance in higher alcohol synthesis (HAS) process for its special nano-alloy structure [10]. However, the Cu phase partly separated from Fe phase during the reaction, which weakened the interaction of iron and copper species. Cu is commonly supposed to offer CO insertion site and facilitate the chemical desorption of H2 and the molecular desorption of CO [13,14]. It is well accepted that Cu-based catalyst modified by Fe species enables the CO chemical adsorption and provides carbon chain growth sites [2,13,14]. Nie et al. [6] reported that Ce promoter improved the formation of FeCx species and donated electron to Fe species in reaction, which leaded to higher CO conversion and lower CH4 selectivity. The importance of the synergistic effect of iron and copper species has been recognized recently, however the effect of ceria nanorod as support of Cu-promoted iron catalyst has rarely been reported.
This paper mainly aims to investigate the effect of supports Mn-modified ceria nanorod in Cu-promoted iron catalyst for HAS. Cu-Fe catalysts supported on Mn-modified ceria nanorod with different amounts of Mn were prepared by the hydrothermal synthesis method and the liquid phase co-reduction method. These catalysts were characterized by N2 physisorption, inductively coupled plasma-atomic emission spectrometer (ICP-AES), field emission scanning electron microscopy (FESEM), transmission electron microscopy (TEM), X-ray diffraction (XRD), temperature programmed reduction (TPR), temperature programmed desorption of adsorbed CO (CO-TPD), X-ray photoelectron spectra (XPS), Mӧssbauer effect spectra (MES), and the reaction performance was performed in a fixed-bed reactor.

2. Results and Discussion

2.1. Textural Properties of the Supports and Catalysts

All the supports display a typical cylinder-like rod, the FESEM images of supports 1.2MnCe, 2.4MnCe and 3.6MnCe are shown in Figure S1, in which the diameter of the support is about 10 nm and length is about 100–300 nm, which is close to the size reported previously [15,16]. The addition of Mn was favorable for the formation of nanoparticles, and Mn-Ce mixed oxide solid solution was possibly formed in this process, and the cylinder-like rod of xMnCe mixed oxides will gradually be replaced by nanoparticles with the continuous increase in Mn content. The result of ICP analysis (Table S1) was commonly in agreement with the theoretical contents, indicating that precursor salts (Fe(NO3)3·9H2O and Cu(NO3)2·3H2O) were reduced completely in the liquid phase co-reduction process. The BET surface area (Table S1) of the catalysts is about 80 m2/g, almost the same with that of the supports, which indicates that the textural properties of the supports are not changed in the catalyst preparation process. TEM image of catalyst CuFe/1.2MnCe (Figure S1) shows that the average size of Cu-Fe bimetallic nanoparticles is about 10–30 nm, which could ensure the good dispersion of metal particles. The XRD patterns (Figure S2a) show that a typical cubic fluoride CeO2 crystal phase (JCPDS#34-0394) exists in the samples. The diffraction peaks of Mn species cannot be observed possibly because of the small amount and the high dispersion of Mn species [15]. The XRD patterns of fresh catalysts are shown in Figure S2b. Peaks at 43.3°, 50.4° and 74.13° could be ascribed to metal copper (JCPDS#04-0836) or/and iron-copper alloy FeCu4 (JCPDS#65-7002) [10]. No peaks of Fe species were observed, possibly because the formed iron species was highly dispersed on the catalysts. The XRD patterns of catalysts after reaction (Figure S3) show that the intensity of the typical cubic fluoride CeO2 crystal phase (JCPDS#34-0394) peaks decreases a significant amount [17], which means that a large part of CeO2 is reduced during the reduction and reaction process. As was shown in Figure S3, diffraction peaks at 26.8°, 29.4° and 48.4° can be ascribed to SiO2 species, which should come from catalysts diluted with 80–100 mesh SiO2.

2.2. Reduction Performance of Catalysts

H2-TPR profiles of fresh supports were shown in Figure 1a. The weak reduction peaks of low temperature at about 200 °C can be assigned to the reduction of MnO2 to Mn3O4 and the major peaks at 380 °C can be deemed to the reduction process of CeO2 to Ce2O3 species. Peaks at ~700 °C can be attributed to the reduction of Ce2O3 to CeO species. H2-TPR profiles of synthesized catalysts were shown in Figure 1b. Reduction peaks located in the range of 100 to 200 °C could be attributed to the reduction process of copper oxide to metal copper (CuO and Cu2O to Cu) [11,18]. The slight reduction peak at about 340 °C corresponds to the surface oxygen reduction of cerium oxide from CeO2 species to Ce2O3 species [17], indicating the transformation of Ce species during the reduction procedure which is in accordance with the results of XRD. The peaks at about 550 °C could be ascribed to the reduction of iron oxide to metal iron (Fe2O3 to Fe3O4, Fe3O4 to FeO and FeO to Fe) [11,18,19]. The iron and copper oxide species were probably formed in the process of centrifugation and passivation, oxidized from metal nanoparticles by air.

2.3. CO Desorption Properties of Catalysts

The desorption peaks of fresh supports (Figure 2) at lower temperature about 150 °C was attributed to the molecular CO state desorption [20]. The peak at higher temperatures of about 270 °C mainly indicated the existence of new CeO2 active sites, mainly involved in the CO catalytic oxidation reaction [21] which led to the CO2 formation in HAS process. Peaks above 500 °C can be regarded as CO chemical desorption, mainly relative to HAS reaction [20]. Obviously, the intensity of desorption peaks was consistent with the Mn content. This change proved that Mn addition promoted the CO adsorption of supports, which led to the formation of more CO species on the catalyst surface and improved the FTS performance in some senses. The desorption curves of different temperature regions illustrated different CO adsorption state [22]. It is generally accepted that the desorption peaks below 200 °C correspond to the molecular CO state desorption on Cu-Fe nanoparticles, while those of the CO chemical desorption mainly correspond to CO hydrogenation, located above 500 °C [20,23]. It can be seen from Figure 2b that the desorption peaks at above 500 °C shift to a higher temperature when the Mn content increases from 0 to 2.4. The proper addition of Mn is beneficial to the enhancement of the strongly adsorbed CO, causing a higher CO concentration over the catalyst surface for the reason that Mn could transfer electrons to CO via iron oxides, and strengthen Fe-C bond, which hinders the CO chemical desorption, causing that more occurred on the surface of catalysts at higher temperatures [24]. A higher temperature of CO desorption is favorable for carbon monoxide insertion, possibly leading to the formation of more long-chain production [1,2,25]. Interestingly, the desorption peak of CuFe/3.6MnCe shifts to a lower temperature compared to other samples, which may be caused by the strong interaction between Fe and Mn when Mn is over added, hindering the CO chemical desorption.

2.4. Surface Chemical Properties of Catalysts

The XPS spectra of Ce 3d5/2 and 3d3/2 of the fresh supports and used catalysts are shown in Figure 3. The XPS spectra of Ce 3d core level can be resolved into 10 groups according to the literature [22,26,27,28,29], three doublets corresponding to Се4+, and two doublets corresponding to Се3+, which has been labeled in Figure 3. The peaks located at 881.7 (Ce3+), 882.9 (Ce4+), 885.7 (Ce3+), 889.0 (Ce4+), and 897.6 (Ce4+) eV correspond to 3d5/2 while the peaks at 899.1 (Ce3+), 900.8 (Ce4+), 904.3 (Ce3+), 907.4 (Ce4+), and 916.8 (Ce4+) eV correspond to Ce 3d3/2. The concentration of the surface Ce3+, which is related to the oxygen vacancies on the catalysts’ surface can be calculated from Equation (1) [29].
C Ce 3 + = Ce 3 + Ce 3 + + Ce 4 +
After the calculation of peak area for Ce4+ and Ce3+, the Ce3+ concentration of fresh catalyst CuFe/0MnCe and the used catalysts are respectively shown in Table 1.
It can be seen from Table 1 that the Ce3+ concentration of used catalyst CuFe/0MnCe is almost twice as large as that of the fresh support, which is in accordance with the results of XRD and H2-TPR. This manifested as Ce4+ transforming to Ce3+ during the reduction and reaction process, which will possibly affect the CO hydrogenation and the selectivity of products. The oxygen vacancy increased with Mn addition (0–0.6 Mn content), but slightly decreased with the increasing Mn content due to excessive Mn addition (0.6–3.6 Mn content), which will inhibit the transformation of Ce4+ to Ce3+ species. Compared with the Ce 3d XPS spectra of fresh supports, the increments of Ce3+ concentration were larger after Mn addition. Mn species intensified the formation process of oxygen vacancies in some senses. It was well accepted that more oxygen vacancies can enhance CO adsorption on the catalyst surface. This phenomenon may be caused by the gradually strengthened interaction between Fe and Mn species and will improve the behavior of CO chemical adsorption.
The Cu 2p spectra and the enlarged spectra within the BE range of 955–948 eV and 935–930 eV of the used catalysts are shown in Figure 4. The peaks at 952.6 eV and 932.6 eV can be attributed to Cu 2p1/2, and Cu 2p3/2, respectively, while those at 952.6 eV and 932.6 eV could be regarded as Cu0 species, with a weak satellite peak at 946.7 eV which is related to Cu+ species [30,31,32,33]. The constitution of Cu species on the used catalysts’ surface is shown in Table 1. The Cu 2p spectra implies that the copper species are mainly constituted by metal copper and a small amount of Cu+, which is consistent with the results of XRD profiles. It was reported that the Cu0 and Cu+ species are preferential for the synthesis of ethanol because of the synergistic effect between balanced Cu+ and Cu0 sites [34,35]. With Mn addition, the ratio of Cu0/Cu+ increases from 6.49 to 9.76.
The Fe 2p spectra of catalysts after reaction are shown in Figure 5. The two main peaks at about 711.5 eV with a shift toward lower binding energy and 724.5 eV with a shift toward lower binding energy are ascribed to Fe 2p3/2 and Fe 2p1/2, respectively [6,36]. The shift of binding energy is relevant to the electron transformation between Fe and Mn species. Mn species tends to transfer electron to Fe species for electronegativity [37]. Therefore, Mn species act as electronic donating promoters and appropriate Mn content accelerates the CO chemical adsorption, which leads to a higher carbon species concentration on the catalyst surface. The peaks at about 707.1 eV are related to iron carbides and its intensity gradually strengthens with increasing Mn addition, indicating that the addition of Mn is preferential for the formation of surface iron carbides species [38,39,40], which seems to be in disagreement with the bulk results of MES.

2.5. MES Results of Used Catalysts

The MES results of the catalysts after reaction under the operating condition of syngas (H2/CO = 2.0), 260 °C, 3.0 MPa and 1333 h−1 are listed in Table 2. It was well known that iron carbide promoted the CO insertion and the formation of more long-chain products [2]. As shown in Table 2, all the used catalysts are composed of Fe3+, Fe3O4, and χ-Fe5C2. The amount of χ-Fe5C2 increases significantly when Mn is added, indicating that the addition of small amounts of Mn promotes the carburization of iron oxides which is commonly accepted as the active phase for CO hydrogenation. However, the excessive addition will lead to a slightly decrease in χ-Fe5C2 content, possibly because of the gradually strengthened interaction of Fe and Mn. At the same time, the decreasing amount of Fe3O4 indicates that part of the Fe3O4 is transformed into iron carbides during the reaction process, possibly because Mn species can donate more electrons in the reaction process [37,41], which is essential to CO chemical adsorption. Meanwhile, Fe3O4 species are commonly regarded as the active phase of the water–gas shift reaction and can facilitate the formation of CO2. The decreasing trend of Fe3O4 content when Mn addition gradually increases is in accordance with the variation of CO2 selectivity during the catalytic performance evaluation process. On the other hand, the slight increase in Fe3+ content may be caused by the oxidation of iron phases by H2O and CO2 during the reaction process [42,43].

2.6. Catalytic Performance and Stability Test of the Catalysts

The catalytic performance tests of catalysts were performed in a tubular fixed-bed reactor, and the results are shown in Table 3. It can be seen that CO conversion decreases largely from 82.83% to 41.43% when Mn addition increases from 0 to 3.6. In the reaction process, CO conversion is mainly caused by CO hydrogenation and catalytic oxidation [1,2]. Fe3O4 species decreased significantly, from 21.9% to 9.7% after 0.6 Mn addition, which explained the decrease in CO conversion from 82.83% to 73.35%. With the continuous increase in Mn addition from 0.6 to 3.6, the active phase χ-Fe5C2, as shown in Table 2, decreased from 47.7% to 44.3%, while Fe3O4 species as active phase for CO catalytic oxidation decreased from 9.7% to 8.7% [6,44]. An excessive Mn addition (0.6–3.6 Mn content) will inhibit the transformation of Ce4+ to Ce3+ and the formation of oxygen vacancies, finally resulting in the loss of CO chemical adsorption. The peak value of CO conversion over CuFe/0MnCe may be caused by the large amount of Fe3O4 species in the catalyst. With Mn addition, the selectivity of CO2 also displays a decreasing trend for the decrease in Fe3O4 species. Meanwhile the strengthened interaction between Mn-Ce mixed oxide will inhibit the transformation of Ce4+ to Ce3+ and reduce the catalytic oxidation of carbon monoxide by cerium dioxide [45], finally leading to a lower selectivity of CO2.
From the results of the reaction performance tests, it can also be seen that the selectivity of total and higher alcohols (C2-5OH and C6+OH) increases with Mn content, especially the selectivity of alcohol over CuFe/3.6MnCe is almost twice as large as that over the catalyst without Mn addition. Relatively, the selectivity of methanol is much lower than that of C2-5OH and C6+OH (84.41 wt% C2+OH over CuFe/3.6MnCe in alcohols distribution). As discussed above, Cu0 is more likely the active site for alcohols synthesis than Cu+ and plays a vital role in the synthesis of alcohols.
The ability of iron carbide to facilitate the CO insertion and the formation of more long-chain products was well known [2]. The excessive addition of Mn will lead to a decrease in the content of iron carbides [45], which will result in a decrease in selectivity of long chain products. On the other hand, the strengthened interaction between Mn and Ce will inhibit the reduction of Ce species, finally reduce the formation of CO2, indirectly increase the selectivity of hydrocarbons (HC) and alcohols. Although the excess addition of Mn will reduce the adsorption amount of CO and decrease the catalyst reactivity, but the addition of Mn could facilitate the carbon chain growth and on the other hand promote the possibility of methanol homologation reaction which will facilitate the conversion of methanol to ethanol and further decrease the content of methanol in the products [11]. For the combination effect of these factors, the selectivity of total alcohols, C2-5OH, C6+OH and hydrocarbons all increase with the addition of Mn.
The reaction stability tests of catalysts were also carried out in syngas (H2/CO = 2.0) at 260 °C and 3.0 MPa with a space velocity of 1333 h−1 for about 140 h. The test outcomes of five catalysts are shown in Figure 6. It can be seen that there is no obvious decrease during the tests and the volatility of CO conversion is about 3%. With Mn addition, CO conversion decreased in sequence: CuFe/0MnCe > CuFe/0.6MnCe > CuFe/1.2MnCe > 6CuFe/2.4MnCe > CuFe/3.6MnCe, and all catalysts maintained stability well.

3. Experimental

3.1. Preparation of Supports and Catalysts

Certain amounts of cerium acetate (12 mmol) and manganese acetate (x mmol) were dissolved in deionized water under magnetic stirring, then mixed with NaOH aqueous solution (8.33 mol/L) and stirred for 0.5 h. The mixed solution then was transferred into two Teflon-lined stainless-steel autoclaves (150 mL) and hydrothermally treated at 100 °C for 24 h to get Mn-modified ceria nanorods (xMnCe, x = 0, 0.6, 1.2, 2.4, 3.6) [15,16,46,47]. The formed slurry was separated by centrifugation and washed with deionized water and ethanol for several times until the pH value of the filter liquor equaled to 7 ± 0.1, then the solid was dried at 110 °C for 12 h.
The prepared support was then dispersed in deionized water, and certain amounts of cupric nitrate (24 mmol) and ferric nitrate (8 mmol) were dissolved in the support-contained deionized water. N2 was flowed into the slurry to remove the dissolved oxygen under vigorous stirring for 1 h. An aqueous solution of NaBH4 (6.8 mol/L) was added into the nitrate solution dropwise under the protection of N2 [10,46,47]. After the reductant addition, the Cu2+ and Fe3+ species were co-reduced into elemental metal, loaded on prepared support (CuFe/xMnCe, x = 0, 0.6, 1.2, 2.4, 3.6). The container of solution was put in an ice-water bath during the whole preparation procedure. The formed slurry was separated by centrifugation and washed with deionized water and methanol several times until the pH value of the filter liquor was equal to 7 ± 0.1.

3.2. Catalyst Characterization

The BET surface area of fresh catalysts (0.2 g and 80–100 mesh) was measured by N2 physisorption at −196 °C by physical adsorption instrument (ASAP 2020 model of America Micromeritics Company, Norcross, GA, USA). The ICP-AES of the catalysts was carried out on Agilent 725 inductively coupled plasma-atomic emission spectrometer (America, Agilent Technologies Inc., Santa Clara, CA, USA). The instrument is equipped with a self-excited RF generator for 40.68 MH and CCD solid state detector. XRD was carried out on a diffraction instrument (D/Max 2550 model of Japanese Rigaku Company, Tokyo, Japan) with Cu Kα radiation (λ = 1.5418 Å), and operated at 40 kV and 40 mA. H2-TPR of samples (50 mg and 80–100 mesh) was carried out by temperature programmed desorption apparatus (Autochem 2920 model of America Micromeritics Company, Norcross, GA, USA) with a thermal conductivity detector (TCD) (temperature from 50 to 800 °C with a rising rate 10 °C/min). CO-TPD was conducted in the same equipment of H2-TPR (temperature from 50 to 1000 °C with a rising rate 10 °C/min). XPS was performed on a VG electron spectrometer (ESCALAB 250Xi model of America Thermo Fisher Scientific Company, Waltham, MA, USA) operated in a constant pass energy mode. The binding energies (BEs) was calibrated by carbonaceous C 1s line (284.6 eV). FESEM images of samples were performed on a Nova NanoSEM 450 microscopy at 3 kV (America Thermoz Fisher Scientific Company, Waltham, MA, USA). TEM images of samples were obtained on a JEOL Model 2100F electron microscopy at 200 kV (Japanese JEOL Company, Tokyo, Japan). The MES of the used catalysts was performed on an MR-351 constant-acceleration Mössbauer spectrometer at room temperature (25 mCi 57Co in a Pd matrix model of German FAST Company, Berlin, Germany).

3.3. Catalytic Performance Evaluation

The reaction performance test of catalysts was performed in a fixed-bed reactor (12 mm i.d.) with 1.0 g fresh catalysts. Fresh catalysts were mixed with 2.0 g SiO2 (80–100 mesh) to eliminate hot spots. The catalysts were reduced in synthesis gas (H2/CO = 2.0) with a space velocity (Sv) of 2400 h−1, with a syngas (H2/CO = 2.0) space velocity of 1333 h−1. The tail gas was on-line analyzed by one Agilent 7890A gas chromatograph equipped with a 5A molecular sieve column for the separation of inorganic gas (CO, CO2, N2, and H2) and a HP-AL/S capillary column for the HC separation (C1–C6 hydrocarbons), then detected by TCD and FID, respectively. Liquid products were collected after stable reaction for 24 h, and then analyzed off-line by another Agilent 7890A gas chromatograph equipped with a HP-5 capillary column for oil phase separation (alkanes, alkenes and alcohols) and DB-WAX for the water phase separation (alcohols). The products of oil phase and water phase were prepared by Agilent 7683B injector and then detected by FID, respectively.
The CO conversation ( X C O ) and selectivity of products ( S i ) were calculated by Equations (S1) and (S2), respectively. The selectivity of C5+ can be calculated by summing up all the selectivity values of hydrocarbons with a carbon chain longer than 5 which was calculated by Equation (S2).
The mass balance was evaluated by calculating the weight ratio of products and reactants after each steady-state reaction period (24 h) and the values were all in the range of 95% to 105%.

4. Conclusions

A series of CuFe/xMnCe catalysts were prepared by the liquid phase co-reduction method and tested in a fixed-bed reactor for higher alcohol synthesis. Cu-Fe nanoparticles were loaded on the surface of Mn-modified cerium dioxide prepared through hydrothermal synthesis method. Catalysts processed the propriety of low temperature reduction well. With the addition of Mn in the ceria nanorod, the capacity of strongly adsorbed CO was enhanced, causing a higher CO concentration over the catalyst surface favorable for higher alcohol synthesis. The transformation of Ce4+ to Ce3+ was inhibited during the reaction process because of the strengthened interaction between Fe and Mn, which will result in the inhibition of CO chemical adsorption. Mn promoter is preferential for the formation of surface iron carbides species, which facilitate the carbon monoxide insertion and carbon chain growth. Cu species play a vital role in higher alcohol synthesis, and Cu0 is more likely the active site for alcohols synthesis. The Cu0/Cu+ ratio increased with the increase in Mn content, which was beneficial to the formation of alcohol products. The highest selectivity over Mn-modified ceria nanorod supported Cu-Fe catalysts is 25.56 wt% with Mn addition 3.6 and CO conversion 41.43%.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/10/10/1124/s1, Figure S1: FESEM images of Mn-Ce mixed oxides nanorods: (a) 1.2MnCe, (b) 2.4MnCe, (c) 3.6MnCe, and TEM image of fresh catalyst: (d) CuFe/2.4MnCe, Figure S2: XRD patterns: (a) xMnCe, (b) CuFe/xMnCe as synthesized, Figure S3: XRD patterns of catalysts CuFe/xMnCe after reaction, Table S1: BET and ICP results of fresh catalysts and the comparison with experimental values.

Author Contributions

Y.X.: Conceptualization, Methodology, Investigation, Writing-original draft, Visualization. H.M.: Conceptualization, Methodology. H.Z.: Writing—review and editing. W.Q.: Writing—review and editing, Supervision, Funding acquisition. Q.S.: Funding acquisition. W.Y.: Writing—review and editing, Supervision, Funding acquisition. D.C.: Writing—Review and Editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China, grant number 21706068, the National High Technology Research and Development Plan of China (863 plan), grant number 2011AA05A204, and the Fundamental Research Funds for the Central Universities, grant number 222202017013.

Conflicts of Interest

No conflict of interest exists in the submission of this manuscript and this manuscript is approved by all authors for publication.

References

  1. Ao, M.; Pham, G.H.; Sunarso, J.; Tade, M.O.; Liu, S. Active centers of catalysts for higher alcohol synthesis from syngas: A review. ACS Catal. 2018, 8, 7025–7050. [Google Scholar] [CrossRef]
  2. Luk, H.T.; Mondelli, C.; Ferré, D.C.; Stewart, J.A.; Pérez-Ramírez, J. Status and prospects in higher alcohols synthesis from syngas. Chem. Soc. Rev. 2017, 46, 1358–1426. [Google Scholar] [CrossRef] [PubMed]
  3. Zaman, S.; Smith, K.J. A review of molybdenum catalysts for synthesis gas conversion to alcohols: Catalysts, mechanisms and kinetics. Catal. Rev. Sci. Eng. 2012, 54, 41–132. [Google Scholar] [CrossRef]
  4. Mills, G.A. Status and future opportunities for conversion of synthesis gas to liquid fuels. Fuel 1994, 73, 1243–1279. [Google Scholar] [CrossRef]
  5. Dorokhov, V.S.; Kamorin, M.A.; Rozhdestvenskaya, N.N.; Kogan, V.M. Synthesis and conversion of alcohols over modified transition metal sulphides. C. R. Chim. 2016, 19, 1184–1193. [Google Scholar] [CrossRef]
  6. Nie, C.; Zhang, H.; Ma, H.; Qian, W.; Sun, Q.; Ying, W. Effects of Ce addition on Fe-Cu catalyst for Fischer–Tropsch synthesis. Catal. Lett. 2019, 149, 1375–1382. [Google Scholar] [CrossRef]
  7. Smith, K.; Anderson, R. A chain growth scheme for the higher alcohols synthesis. J. Catal. 1984, 85, 428–436. [Google Scholar] [CrossRef]
  8. Boz, I.; Sahibzada, M.; Metcalfe, I. Kinetics of the higher alcohol synthesis over a K-promoted CuO/ZnO/Al2O3 catalyst. Ind. Eng. Chem. Res. 1994, 33, 2021–2028. [Google Scholar] [CrossRef]
  9. Camposmartin, J.M.; Guerreroruiz, A.; Fierro, J.L.G. Structural and surface properties of CuO-ZnO-Cr2O3 catalysts and their relationship with selectivity to higher alcohol synthesis. J. Catal. 1995, 156, 208–218. [Google Scholar] [CrossRef]
  10. Xiao, K.; Qi, X.; Bao, Z.; Wang, X.; Zhong, L.; Fang, K.; Lin, M.; Sun, Y. CuFe, CuCo and CuNi nanoparticles as catalysts for higher alcohol synthesis from syngas: A comparative study. Catal. Sci. Technol. 2013, 3, 1591–1602. [Google Scholar] [CrossRef]
  11. Qian, W.; Wang, H.; Xu, Y.; Yang, X.; Zhai, G.; Zhang, H.; Ma, H.; Sun, Q.; Ying, W. In situ DRIFTS study of homologous reaction of methanol and higher alcohols synthesis over Mn promoted Cu-Fe catalysts. Ind. Eng. Chem. Res. 2019, 58, 6288–6297. [Google Scholar] [CrossRef]
  12. Ding, M.; Qiu, M.; Liu, J.; Li, Y.; Wang, T.; Ma, L.; Wu, C. Influence of manganese promoter on co-precipitated Fe-Cu based catalysts for higher alcohols synthesis. Fuel 2013, 109, 21–27. [Google Scholar] [CrossRef]
  13. Maksimov, Y.; Suzdalev, I.; Arents, R.; Loktev, S. Study of nucleation of carbide phase on defective sections of promoted iron catalyst by gamma-resonance spectroscopy method. Russ. Chem. Bull. 1974, 23, 1542–1544. [Google Scholar] [CrossRef]
  14. Hindermann, J.; Razzaghi, A.; Breault, R.; Kieffer, R.; Kiennemann, A. Orientation towards alcohols in CO, H2 reactions on some iron catalysts. React. Kinet. Catal. Lett. 1984, 26, 221–226. [Google Scholar] [CrossRef]
  15. Liao, Y.; Fu, M.; Chen, L.; Wu, J.; Huang, B.; Ye, D. Catalytic oxidation of toluene over nanorod-structured Mn–Ce mixed oxides. Catal. Today 2013, 216, 220–228. [Google Scholar] [CrossRef]
  16. Mai, H.X.; Sun, L.D.; Zhang, Y.W.; Si, R.; Feng, W.; Zhang, H.P.; Liu, H.C.; Yan, C.H. Shape-selective synthesis and oxygen storage behavior of ceria nanopolyhedra, nanorods, and nanocubes. J. Phys. Chem. B 2005, 109, 24380–24385. [Google Scholar] [CrossRef] [PubMed]
  17. Hu, Z.; Liu, X.; Meng, D.; Guo, Y.; Guo, Y.; Lu, G. Effect of ceria crystal plane on the physicochemical and catalytic properties of Pd/ceria for CO and propane oxidation. ACS Catal. 2016, 6, 2265–2279. [Google Scholar] [CrossRef]
  18. Das, S.K.; Mohanty, P.; Majhi, S.; Pant, K.K. CO-hydrogenation over silica supported iron based catalysts: Influence of potassium loading. Appl. Energy 2013, 111, 267–276. [Google Scholar] [CrossRef]
  19. Pengnarapat, S.; Ai, P.; Reubroycharoen, P.; Vitidsant, T.; Yoneyama, Y.; Tsubaki, N. Active Fischer-Tropsch synthesis Fe-Cu-K/SiO2 catalysts prepared by autocombustion method without a reduction step. J. Energy Chem. 2018, 27, 432–438. [Google Scholar] [CrossRef]
  20. Suo, H.; Wang, S.; Zhang, C.; Xu, J.; Wu, B.; Yang, Y.; Xiang, H.; Li, Y. Chemical and structural effects of silica in iron-based Fischer–Tropsch synthesis catalysts. J. Catal. 2012, 286, 111–123. [Google Scholar] [CrossRef]
  21. Zou, Q.; Zhao, Y.; Jin, X.; Fang, J.; Li, D.; Li, K.; Lu, J.; Luo, Y. Ceria-nano supported copper oxide catalysts for CO preferential oxidation: Importance of oxygen species and metal-support interaction. Appl. Surf. Sci. 2019, 494, 1166–1176. [Google Scholar] [CrossRef]
  22. Liu, C.; Zhou, J.; Ma, H.; Qian, W.; Zhang, H.; Ying, W. Antisintering and high-activity Ni catalyst supported on mesoporous silica incorporated by Ce/Zr for CO methanation. Ind. Eng. Chem. Res. 2018, 57, 14406–14416. [Google Scholar] [CrossRef]
  23. Benziger, J.; Madix, R.J.S. The effects of carbon, oxygen, sulfur and potassium adlayers on CO and H2 adsorption on Fe (100). Surf. Sci. 1980, 94, 119–153. [Google Scholar] [CrossRef]
  24. Zhu, H. Supported rhodium and supported aqueous-phase catalyst, and supported rhodium catalyst modified with water-soluble TPPTS ligands. Appl. Catal. A 2003, 245, 111–117. [Google Scholar]
  25. Xu, D.; Zhang, H.; Ma, H.; Qian, W.; Ying, W. Effect of Ce promoter on Rh-Fe/TiO2 catalysts for ethanol synthesis from syngas. Catal. Commun. 2017, 98, 90–93. [Google Scholar] [CrossRef]
  26. Bêche, E.; Charvin, P.; Perarnau, D.; Abanades, S.; Flamant, G. Ce 3d XPS investigation of cerium oxides and mixed cerium oxide (CexTiyOz). Surf. Interface Anal. 2008, 40, 264–267. [Google Scholar] [CrossRef]
  27. Romeo, M.; Bak, K.; El Fallah, J.; Le Normand, F.; Hilaire, L.J.S.; Analysis, I. XPS study of the reduction of cerium dioxide. Surf. Interface Anal. 1993, 20, 508–512. [Google Scholar] [CrossRef]
  28. Chang, L.H.; Sasirekha, N.; Chen, Y.W.; Wang, W.J. Preferential oxidation of CO in H2 stream over Au/MnO2-CeO2 catalysts. Ind. Eng. Chem. Res. 2006, 45, 4927–4935. [Google Scholar] [CrossRef]
  29. Stadnichenko, A.I.; Murav’ev, V.V.; Svetlichnyi, V.A.; Boronin, A.I. Platinum state in highly active Pt/CeO2 catalysts from the X-ray photoelectron spectroscopy data. J. Struct. Chem. 2017, 58, 1152–1159. [Google Scholar] [CrossRef]
  30. Chen, L.; Guo, P.; Qiao, M.; Yan, S.; Li, H.; Shen, W.; Xu, H.; Fan, K. Cu/SiO2 catalysts prepared by the ammonia-evaporation method: Texture, structure, and catalytic performance in hydrogenation of dimethyl oxalate to ethylene glycol. J. Catal. 2008, 257, 172–180. [Google Scholar] [CrossRef]
  31. Gervasini, A.; Manzoli, M.; Martra, G.; Ponti, A.; Ravasio, N.; Sordelli, L.; Zaccheria, F. Dependence of copper species on the nature of the support for dispersed CuO catalysts. J. Phys. Chem. B 2006, 110, 7851–7861. [Google Scholar] [CrossRef] [PubMed]
  32. Liu, Y.; Zuo, Z.; Li, C.; Deng, X.; Huang, W. Effect of preparation method on CuZnAl catalysts for ethanol synthesis from syngas. Appl. Surf. Sci. 2015, 356, 124–127. [Google Scholar] [CrossRef]
  33. Li, M.; Jiao, L.; Nawaz, M.A.; Cheng, L.; Meng, C.; Yang, T.; Tariq, M.; Liu, D. A one-step synthesis method of durene directly from syngas using integrated catalyst of Cu/ZnO/Al2O3 and Co-Nb/HZSM-5. Chem. Eng. Sci. 2019, 200, 103–112. [Google Scholar] [CrossRef]
  34. Wen, C.; Cui, Y.; Chen, X.; Zong, B.; Dai, W.-L. Reaction temperature controlled selective hydrogenation of dimethyl oxalate to methyl glycolate and ethylene glycol over copper-hydroxyapatite catalysts. Appl. Catal. B 2015, 162, 483–493. [Google Scholar] [CrossRef]
  35. Zhu, Y.; Kong, X.; Zhu, S.; Dong, F.; Zheng, H.; Zhu, Y.; Li, Y. Construction of Cu/ZrO2/Al2O3 composites for ethanol synthesis: Synergies of ternary sites for cascade reaction. Appl. Catal. B 2015, 166–167, 551–559. [Google Scholar] [CrossRef]
  36. Yamashita, T.; Hayes, P. Analysis of XPS spectra of Fe2+ and Fe3+ ions in oxide materials. Appl. Surf. Sci. 2008, 254, 2441–2449. [Google Scholar] [CrossRef]
  37. Qin, S.; Zhang, C.; Xu, J.; Yang, Y.; Xiang, H.; Li, Y. Fe–Mo interactions and their influence on Fischer–Tropsch synthesis performance. Appl. Catal. A 2011, 392, 118–126. [Google Scholar] [CrossRef]
  38. Bonnet, F.; Ropital, F.; Lecour, P.; Espinat, D.; Huiban, Y.; Gengembre, L.; Berthier, Y.; Marcus, P. Study of the oxide/carbide transition on iron surfaces during catalytic coke formation. Surf. Interface Anal. 2002, 34, 418–422. [Google Scholar] [CrossRef]
  39. Lin, T.C.; Seshadri, G.; Kelber, J.A. A consistent method for quantitative XPS peak analysis of thin oxide films on clean polycrystalline iron surfaces. Appl. Surf. Sci. 1997, 119, 83–92. [Google Scholar] [CrossRef]
  40. Graat, P.C.J.; Somers, M.A.J. Simultaneous determination of composition and thickness of thin iron-oxide films from XPS Fe 2p spectra. Appl. Surf. Sci. 1996, 100, 36–40. [Google Scholar] [CrossRef]
  41. Qin, S.; Zhang, C.; Xu, J.; Wu, B.; Xiang, H.; Li, Y. Effect of Mo addition on precipitated Fe catalysts for Fischer–Tropsch synthesis. J. Mol. Catal. A Chem. 2009, 304, 128–134. [Google Scholar] [CrossRef]
  42. Ding, M.; Yang, Y.; Xu, J.; Tao, Z.; Wang, H.; Wang, H.; Xiang, H.; Li, Y. Effect of reduction pressure on precipitated potassium promoted iron–manganese catalyst for Fischer–Tropsch synthesis. Appl. Catal. A 2008, 345, 176–184. [Google Scholar] [CrossRef]
  43. Wang, H.; Yang, Y.; Xu, J.; Wang, H.; Ding, M.; Li, Y. Study of bimetallic interactions and promoter effects of FeZn, FeMn and FeCr Fischer–Tropsch synthesis catalysts. J. Mol. Catal. A Chem. 2010, 326, 29–40. [Google Scholar] [CrossRef]
  44. Li, X.; Zhong, B.; Peng, S.; Wang, Q. Fischer-Tropsch synthesis on Fe-Mn ultrafine catalysts. Catal. Lett. 1994, 23, 245–250. [Google Scholar] [CrossRef]
  45. Razzaq, R.; Zhu, H.; Jiang, L.; Muhammad, U.; Li, C.; Zhang, S. Catalytic methanation of CO and CO2 in coke ovengas over Ni–Co/ZrO2–CeO2. Ind. Eng. Chem. Res. 2013, 52, 2247–2256. [Google Scholar] [CrossRef]
  46. Singh, S.K.; Singh, A.K.; Aranishi, K.; Xu, Q. Noble-metal-free bimetallic nanoparticle-catalyzed selective hydrogen generation from hydrous hydrazine for chemical hydrogen storage. J. Am. Chem. Soc. 2011, 133, 19638–19641. [Google Scholar] [CrossRef]
  47. Bao, H.J.; Li, Y.N.; Liu, L.; Ai, Y.J.; Zhou, J.J.; Qi, L.; Jiang, R.H.; Hu, Z.A.; Wang, J.T.; Sun, H.B.; et al. Ultrafine FeCu alloy nanoparticles magnetically immobilized in amine-rich silica spheres for dehalogenation-proof hydrogenation of nitroarenes. Chem. Eur. J. 2018, 24, 14418–14424. [Google Scholar] [CrossRef]
Figure 1. H2-TPR profiles: (a) fresh supports, (b) synthesized catalysts.
Figure 1. H2-TPR profiles: (a) fresh supports, (b) synthesized catalysts.
Catalysts 10 01124 g001
Figure 2. CO-TPD profiles: (a) supports, (b) fresh catalysts.
Figure 2. CO-TPD profiles: (a) supports, (b) fresh catalysts.
Catalysts 10 01124 g002
Figure 3. Ce 3d XPS spectra: (a) fresh supports, (b) used catalysts.
Figure 3. Ce 3d XPS spectra: (a) fresh supports, (b) used catalysts.
Catalysts 10 01124 g003
Figure 4. Cu 2p XPS spectra of catalysts after reaction: (a) Cu 2p, (b) Cu 2p1/2, (c) Cu 2p3/2.
Figure 4. Cu 2p XPS spectra of catalysts after reaction: (a) Cu 2p, (b) Cu 2p1/2, (c) Cu 2p3/2.
Catalysts 10 01124 g004aCatalysts 10 01124 g004b
Figure 5. Fe 2p XPS spectra of catalysts after reaction.
Figure 5. Fe 2p XPS spectra of catalysts after reaction.
Catalysts 10 01124 g005
Figure 6. Catalytic stability tests of CuFe/xMnCe catalysts.
Figure 6. Catalytic stability tests of CuFe/xMnCe catalysts.
Catalysts 10 01124 g006
Table 1. Ce and Cu compositions on the surface of supports and used catalysts.
Table 1. Ce and Cu compositions on the surface of supports and used catalysts.
SupportsCe3+ Concentration (%)Used CatalystsCe3+ Concentration (%)Value of Cu0/Cu+Increment 1 of Ce3+ Concentration (%)
0MnCe35.89CuFe/0MnCe64.256.4928.36
0.6MnCe33.26CuFe/0.6MnCe65.166.5431.90
1.2MnCe31.49CuFe/1.2MnCe64.937.5733.44
2.4MnCe29.38CuFe/2.4MnCe64.898.0335.51
3.6MnCe27.58CuFe/3.6MnCe64.119.7636.53
1 Increment = Ce3+ Concentration of used catalysts—Ce3+ Concentration of supports.
Table 2. MES results of catalysts after reaction.
Table 2. MES results of catalysts after reaction.
AssignmentsIron Phase Composition of Catalysts (%)
Fe3+ (spm)Fe3O4χ-Fe5C2
CuFe/0MnCe39.221.938.9
CuFe/0.6MnCe42.69.747.7
CuFe/1.2MnCe44.29.546.3
CuFe/2.4MnCe45.49.245.4
CuFe/3.6MnCe47.08.744.3
Table 3. HAS evaluation results of CuFe/xMnCe catalysts.
Table 3. HAS evaluation results of CuFe/xMnCe catalysts.
SamplesXCO
(%)
Selectivity (wt%)Alcohols Distribution (wt%)
CO2ROHHCMeOHC2–5OHC6+OH
CuFe/0MnCe82.8350.9613.7735.2721.6257.6420.74
CuFe/0.6MnCe73.3546.4317.2036.3721.0257.9021.08
CuFe/1.2MnCe59.0941.1920.1338.6819.8958.2321.88
CuFe/2.4MnCe46.1236.8322.5540.6217.5558.6323.82
CuFe/3.6MnCe41.4334.2525.5640.1915.5959.0425.37

Share and Cite

MDPI and ACS Style

Xu, Y.; Ma, H.; Zhang, H.; Qian, W.; Sun, Q.; Ying, W.; Chen, D. Cu-Promoted Iron Catalysts Supported on Nanorod-Structured Mn-Ce Mixed Oxides for Higher Alcohol Synthesis from Syngas. Catalysts 2020, 10, 1124. https://doi.org/10.3390/catal10101124

AMA Style

Xu Y, Ma H, Zhang H, Qian W, Sun Q, Ying W, Chen D. Cu-Promoted Iron Catalysts Supported on Nanorod-Structured Mn-Ce Mixed Oxides for Higher Alcohol Synthesis from Syngas. Catalysts. 2020; 10(10):1124. https://doi.org/10.3390/catal10101124

Chicago/Turabian Style

Xu, Yanbo, Hongfang Ma, Haitao Zhang, Weixin Qian, Qiwen Sun, Weiyong Ying, and De Chen. 2020. "Cu-Promoted Iron Catalysts Supported on Nanorod-Structured Mn-Ce Mixed Oxides for Higher Alcohol Synthesis from Syngas" Catalysts 10, no. 10: 1124. https://doi.org/10.3390/catal10101124

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