Yaoya
Luo
ab,
Sen
Wang
*a,
Shujia
Guo
ab,
Kai
Yuan
ab,
Hao
Wang
a,
Mei
Dong
a,
Zhangfeng
Qin
*a,
Weibin
Fan
a and
Jianguo
Wang
*b
aState Key Laboratory of Coal Conversion, Institute of Coal Chemistry, Chinese Academy of Sciences, P. O. Box 165, Taiyuan, Shanxi 030001, PR China. E-mail: wangsen@sxicc.ac.cn; qzhf@sxicc.ac.cn; Fax: +86 351 4041153; Tel: +86 351 4046092
bUniversity of the Chinese Academy of Sciences, Beijing 100049, PR China. E-mail: iccjgw@sxicc.ac.cn
First published on 2nd November 2020
The surface oxygen vacancy concentration on ZnCeZrO composite oxides was finely tuned through modulating the preparation method and altering the precursor calcination temperature; the effect of the surface oxygen vacancy concentration on the performance of bifunctional ZnCeZrO/SAPO-34 catalysts in the direct conversion of syngas into light olefins (STO) was then investigated. The results indicate that the surface oxygen vacancy concentration of Zn0.5CeZrOx composite oxides can be markedly elevated by preparation through a sol–gel method with glucose as the complexing agent and calcination at 500 °C; a higher surface oxygen vacancy concentration leads to a higher space–time yield of methanol for syngas conversion over the Zn0.5CeZrOx oxide. When combined with SAPO-34 molecular sieves, the bifunctional Zn0.5CeZrOx/SAPO-34 catalyst with high surface oxygen vacancy concentration also exhibits a high space–time yield of light olefins (ethene to butenes) in the synthesis of olefins directly from syngas. With the help of in situ DRIFTS, it can be concluded that the surface oxygen vacancies on the ZnCeZrO oxide play an important role in the catalytic conversion of syngas. The abundant surface oxygen vacancies on Zn0.5CeZrOx can improve the formation of methanol-related intermediates from the syngas over the Zn0.5CeZrOx moiety, promote the evolution of these intermediates into the olefin products over the SAPO-34 moiety, and then enhance the overall capacity of the bifunctional Zn0.5CeZrOx/SAPO-34 composite catalyst in STO.
For STO over bifunctional OX–ZEO catalysts, it is now widely accepted that CO molecules are first adsorbed and activated on the surface oxygen vacancies of metal oxides, which interact with the adjacent active H* species generated from the dissociation of H2, forming formaldehyde (HCO*), formate (HCOO*) and methoxy (H3CO*) intermediates as well as methanol through successive hydrogenation reactions.14 These methanol-related intermediates or methanol molecules migrate or diffuse quickly onto the acid sites of zeolites, producing light olefins through the well-known hydrocarbon pool (HCP) mechanism.15 Undoubtedly, an efficient STO process via the methanol intermediated route needs a bifunctional OX–ZEO catalyst that can effectively couple these two steps, where the surface oxygen vacancies of metal oxides may play a very important role in the formation of methanol-related intermediates and the subsequent evolution of these intermediates into olefins.13
In this work, the surface oxygen vacancy concentration on ZnCeZrO composite oxides was finely tuned through modulating the sol–gel preparation method by varying the complexing agents and altering the precursor calcination temperature; the effect of the surface oxygen vacancy concentration on the performance of bifunctional ZnCeZrO/SAPO-34 catalysts in the direct conversion of syngas into light olefins (STO) was then investigated.
To evaluate the influence of calcination temperature on the surface properties of ZnCeZrO composite oxides, the ZnCeZrO oxide precursor (prepared with glucose as the complexing agent and pre-calcined at 300 °C for 1 h) was subjected to calcination at different temperatures (400, 500, 600 and 700 °C) for 3 h in air; the resultant products are denoted as Zn0.5CeZrOx-glucose-T, where T is the calcination temperature in °C.
SAPO-34 molecular sieves were synthesized by a hydrothermal method with silica sol (JN-40), phosphoric acid (H3PO4), pseudo-boehmite (Al2O3) and tetraethyl ammonium hydroxide (TEAOH). The synthesis gel with the composition 2.0TEAOH:0.05SiO2:1.0Al2O3:1.0P2O5:70H2O was sealed into a Teflon-lined stainless steel autoclave and then crystallized under rotation conditions (15 rpm) at 200 °C for 20 h. H-SAPO-34 molecular sieves were attained by directly calcining the as-synthesized SAPO-34 sample at 550 °C for 10 h in air.
The ZnCeZrO composite oxide and H-SAPO-34 powder with the same mass were then mixed, compressed into tablets and crushed and sieved to 20–40 mesh, to obtain the bifunctional ZnCeZrO/SAPO-34 catalysts for evaluation in the conversion of syngas into light olefins (STO). It should be noted that the composite catalyst here was prepared by uniformly mixing the powdered ZnCeZrO oxide and the powdered SAPO-34 molecular sieves with a mass ratio of 1 (unless specifically indicated). Although the bifunctional catalysts could be realized through various ways for the assembly of two catalyst components such as integrated synthesis, powder mixing, granule stacking and dual bed filling, it was demonstrated recently that the powder mixing of metal oxide and zeolite moieties in general could achieve an intimate contact between the two components, facilitating the transfer of the methanol-related intermediates formed on the metal oxide moiety to the zeolite moiety, leading to a high yield of light olefins.16,17
The field emission-scanning electron microscopy (FE-SEM) images were taken on a JEOL JSM-7001F microscope. High-resolution transmission electron microscopy (HR-TEM) and TEM images were obtained on a field emission-transmission electron microscope (JEM-2100F, JEOL). The average size of catalyst particles was estimated by counting more than 100 particles shown in the TEM images. Scanning transmission electron microscopy (STEM) images and elemental distributions were acquired at 200 kV using a JEOL JEM-2100F microscope equipped with an energy-dispersive X-ray spectroscopy (EDX) detector.
The textural properties were measured by N2 sorption on a Micromeritics TriStar II 3020 instrument at −196 °C. Prior to the measurement, the catalyst samples were degassed at 250 °C for 8 h. The total surface area was obtained from the adsorption branch isotherm in the relative pressure range of 0.05–0.25 by the BET method, the pore volume was calculated from the desorption isotherm by the t-plot method, whilst the pore size distribution was determined by the BJH method.
The Raman spectra were collected on a LabRAM HR Evolution Raman spectrometer equipped with a 532 nm Ar+ laser.
The X-ray photoelectron spectra (XPS) of O (1s) and Ce (3d) were measured at 5 × 10−7 Pa on an AXIS ULTRA DLD instrument with an Al Kα monochromator X-ray source (hν = 1486.6 eV) and calibrated with the binding energy of carbonaceous deposit (C 1s, 284.6 eV).
The temperature-programmed desorption of CO or NH3 (CO-TPD or NH3-TPD) was carried out on a Micromeritics AutoChem II 2920 apparatus. First, about 0.1 g catalyst was loaded into the sample tube and pretreated at 300 °C for 1 h in a He flow; it was then cooled to room temperature to allow saturated adsorption of CO or NH3. After that, the catalyst sample was swept with a He flow and the CO-TPD or NH3-TPD profiles were then recorded at 50–600 °C with a heating rate of 10 °C min−1, through measuring the quantity of desorbed CO or NH3 with a TCD.
The time-dependent diffuse reflectance infrared Fourier transform (DRIFT) spectra were recorded on a Bruker Vertex 80 infrared spectrometer equipped with a liquid nitrogen-cooled MCT detector and an in situ high-temperature chamber. First, about 50 mg of catalyst sample was placed in the in situ chamber and pretreated at 400 °C for 2 h in a H2 flow (30 mL min−1). Next, it was purged with an Ar flow (30 mL min−1) for 0.5 h and cooled to the reaction temperature (300 °C) to collect the background spectrum. After that, the catalyst sample was exposed to a syngas flow (CO/H2 = 1/2, 45 mL min−1) at 300 °C; the IR spectra in the range of 4000–1000 cm−1 were then collected every 5 min up to 90 min at a resolution of 4 cm−1 by accumulating 128 scans.
The effluent products were periodically analyzed online by using an Agilent 7890A gas chromatograph (GC) equipped with one TCD and two FIDs and two capillary columns (J&W 127-7031, 30 m × 530 μm × 0.25 μm; Agilent 19095P-S25, 50 m × 530 μm × 15 μm). The conversion of CO (x(CO)) and the selectivity to CO2 (s(CO2)) are calculated as
x(CO) = (n(CO,in) − n(CO,out))/n(CO,in) × 100% |
s(CO2) = n(CO2,out)/(n(CO,in) − n(CO,out)) × 100% |
s(CiHix) = i·n(CiHix)/(∑j·n(CjHjx) + n(CH3OH) + 2n(DME)) × 100% |
s(CH3OH) = n(CH3OH)/(∑j·n(CjHix) + n(CH3OH) + 2n(DME)) × 100% |
s(DME) = 2n(DME)/(∑j·n(CjHjx) + n(CH3OH) + 2n(DME)) × 100% |
The particle sizes estimated from the diffraction peak at 29.1° by using the Scherrer equation as well as the cell parameters determined by the Rietveld refinement of the XRD patterns referenced to the cubic CeZrO4 solid solution are given in Table 1. No diffraction peaks belonging to the ZnO phase are detected in the XRD patterns, suggesting that the Zn species are highly dispersed in the Zn0.5CeZrOx composite oxides. Meanwhile, as shown in Fig. 1(IV) and S2, ESI,† the Zn0.5CeZrOx composite oxides appear as aggregates of small spherical nanoparticles (NPs) with a mean size of around 3–7 nm, in line with the XRD results (Table 1). The STEM-EDX elemental mapping results shown in Fig. 1(V) further illustrate that the Zn, Ce and Zr elements are uniformly dispersed in the Zn0.5CeZrOx composite oxides.
Entry | Sample | Particle size (nm) | Cell parameter (Å) | S BET (m2 g−1) | V micro (cm3 g−1) |
---|---|---|---|---|---|
Note: (a) the particle sizes reported here were estimated from the diffraction peak at 29.1° in the XRD patterns by using the Scherrer equation. (b) The cell parameters were determined by the Rietveld refinement of the XRD patterns referenced to the cubic CeZrO4 solid solution. (c) The surface areas (SBET) and micropore volumes (Vmicro) were derived from the nitrogen sorption isotherms by the BET and t-plot methods, respectively. | |||||
A | Zn0.5CeZrOx-glucose-500 | 4.5 | 10.5476 | 72 | 0.116 |
B | Zn0.5CeZrOx-citric-500 | 4.0 | 10.5518 | 34 | 0.044 |
C | Zn0.5CeZrOx-tartaric-500 | 4.1 | 10.5393 | 37 | 0.044 |
D | Zn0.5CeZrOx-adipic-500 | 4.3 | 10.5343 | 25 | 0.053 |
E | Zn0.5CeZrOx-alanine-500 | 3.4 | 10.5866 | 15 | 0.018 |
A1 | Zn0.5CeZrOx-glucose-400 | 4.0 | 10.5524 | 115 | 0.131 |
A2(A) | Zn0.5CeZrOx-glucose-500 | 4.5 | 10.5476 | 72 | 0.116 |
A3 | Zn0.5CeZrOx-glucose-600 | 5.0 | 10.5474 | 47 | 0.089 |
A4 | Zn0.5CeZrOx-glucose-700 | 6.5 | 10.5437 | 39 | 0.092 |
The textural properties given in Table 1 (as determined from the nitrogen sorption isotherms shown in Fig. S3, ESI†) show that the Zn0.5CeZrOx-glucose composite oxide prepared with glucose as the complexing agent has a much larger surface area (72 m2 g−1) and micropore volume (0.116 cm3 g−1) than the Zn0.5CeZrOx samples synthesized with other complexing agents (14–37 m2 g−1 and 0.018–0.053 cm3 g−1). Meanwhile, for the Zn0.5CeZrOx-glucose composite oxide, as given in Table 1 (Fig. S4, ESI†), with an increase in the calcination temperature from 400 to 700 °C, the surface area decreases from 115 to 39 m2 g−1, whereas the average particle size increases slightly from 4.0 to 6.5 nm (estimated by using the Scherrer equation from the XRD patterns shown in Fig. S5, ESI†); in particular, weak diffraction lines for the ZnO phase are also detected in the XRD patterns of Zn0.5CeZrOx-glucose calcined at high temperature (>600 °C).
As the second component of the bifunctional ZnCeZrO/SAPO-34 catalysts, as shown in Fig. 2, the as-synthesized SAPO-34 molecular sieves show a typical CHA crystal structure (Fig. 2(I)), with a cubic morphology and an average crystal size of about 300 nm (Fig. 2(II)). The surface area and pore volume determined by nitrogen sorption are 442 m2 g−1 and 0.26 cm3 g−1, respectively. Fig. 2(III) shows the NH3-TPD profile of SAPO-34; it was estimated that the densities of weak (peaking at 170 °C) and strong (at 350 °C) acid sites are 0.28 and 0.45 mmol g−1, respectively. After the powder mixing of the Zn0.5CeZrOx composite oxide with the H-SAPO-34 molecular sieve moieties, as shown in Fig. S6 and S7, ESI,† the crystal structure of the two components is well maintained, whereas a moderate decrease in the surface area, pore volume and quantity of acid sites (in particular the strong acid sites) is observed, originating from the intimate interaction between the metal oxide and molecular sieve moieties.
Fig. 2 Structural characterization of SAPO-34. (I) XRD pattern of SAPO-34. (II) SEM image of SAPO-34. (III) NH3-TPD profile of SAPO-34. |
The surface oxygen vacancies of the Zn0.5CeZrOx composite oxides were characterized by Raman spectroscopy and XPS, as shown in Fig. 3 and S8, ESI.† In the Raman spectra (Fig. 3(I)), the peak at 475 cm−1 corresponds to the F2g symmetrical stretching vibration in the cubic fluorite structure of CeO2.20 Another two peaks at around 320 and 630 cm−1 are ascribed to the displacement of O atoms from their ideal fluorite lattice position, as the introduction of Zr in the Zn0.5CeZrOx composite oxide may distort the Ce–O bonds.21 The surface oxygen vacancy concentration of the Zn0.5CeZrOx composite oxides can then be estimated from the intensity of the peak at 630 cm−1,22 as given in Table 2. Obviously, the Zn0.5CeZrOx-glucose composite oxide shows a much higher surface oxygen vacancy concentration than the other samples, suggesting that high concentration of surface oxygen vacancies or defect sites was created in Zn0.5CeZrOx when glucose was used as the complexing agent in the sol–gel preparation.
Moreover, the concentration of surface oxygen vacancies in Zn0.5CeZrOx-glucose is also greatly influenced by the calcination temperature; the Zn0.5CeZrOx-glucose composite calcined at 500 °C has the highest concentration of surface oxygen vacancies (Table 2). An adequate calcination temperature (500 °C) is probably necessary to promote the mutual isomorphous substitution of Zr4+ and Ce4+ as well as the formation of abundant defect sites in the composite oxide; however, a higher calcination temperature (>500 °C) may lead to a decrease in the oxygen storage capacity and the surface oxygen vacancy concentration, due to the sintering and maturation of the Zn0.5CeZrOx composite oxides at high temperature.
Fig. 3(II) shows the O 1s XPS spectra of the Zn0.5CeZrOx composite oxides. The peaks at 529.0, 530.5, 531.6 and 532.0 eV, representing the lattice O species (Olattice), O atoms around vacancies (Ov), O atoms in the surface hydroxyl (–OH) and O atoms in the absorbed water (H2O), respectively,23 are clearly observed. On the basis of the relative intensity of these peaks, the concentration of oxygen at defect sites was then estimated, as given in Table 2.
Entry | Sample | C Ov (%) | C O-defect (%) | C Ce3+ (%) | CH3OH STY (mmol h−1 g−1) | C2=–C4= STY (mmol h−1 g−1) |
---|---|---|---|---|---|---|
Note: (1) the concentration of surface oxygen vacancies (COv) was obtained from the Raman spectra as COv = Ov/Ototal = I630/(I475 + I630) × 100%,22 where I630 and I475 correspond to the intensity of the peaks at 600 and 475 cm−1, respectively. (2) The concentration of oxygen at defect sites (CO-defect) was calculated from the O 1s XPS spectra as CO-defect = IO-defect/(IO-defect + IO-lattice) × 100%,23 where IO-defect and IO-lattice are the intensity of the peaks at 530.5 and 529.5 eV, respectively. (3) The concentration of Ce3+ on the surface (CCe3+) was calculated from the Ce 3d XPS spectra as CCe3+ = (I(u1) + I(v1))/∑(I(ui) + I(vi)) × 100%,25–27 where I(x) represents the signal intensity of u0 (900.9 eV) and v0 (882.5 eV), u1 (903.2 eV) and v1 (884.7 eV), u2 (907.3 eV) and v2 (888.8 eV), and u3 (916.6 eV) and v3 (898.4 eV) in the Ce 3d XPS spectra. (4) The CH3OH STY refers to the space–time yield of methanol for the syngas conversion over the Zn0.5CeZrOx composite oxides, at 300 °C, 1 MPa, and CO/H2 = 1/2, with a GHSV of 5400 mL g−1 h−1, reported at a time on stream (TOS) of 30 h. (5) The C2=–C4= STY is the space–time yield of C2=–C4= olefins on the basis of carbon, for the syngas conversion over the corresponding bifunctional Zn0.5CeZrOx/SAPO-34 catalyst with an oxide/zeolite mass ratio of 1, at 300 °C, 1 MPa and CO/H2 = 1/2, with a GHSV of 5400 mL g−1 h−1, reported at a TOS of 30 h. (6) The CH3OH STYs for syngas conversion over the Zn0.5CeZrOx-glucose oxides calcined at 400, 600 and 700 °C were not determined. | ||||||
A | Zn0.5CeZrOx-glucose-500 | 23.8 | 29.7 | 30.0 | 0.24 | 0.40 |
B | Zn0.5CeZrOx-citric-500 | 22.2 | 26.9 | 28.9 | 0.19 | 0.25 |
C | Zn0.5CeZrOx-tartaric-500 | 20.9 | 25.1 | 27.0 | 0.18 | 0.20 |
D | Zn0.5CeZrOx-adipic-500 | 19.3 | 24.4 | 25.0 | 0.15 | 0.17 |
E | Zn0.5CeZrOx-alanine-500 | 18.9 | 24.2 | 24.3 | 0.13 | 0.13 |
A1 | Zn0.5CeZrOx-glucose-400 | 21.6 | 21.6 | 21.7 | n.d. | 0.37 |
A2(A) | Zn0.5CeZrOx-glucose-500 | 23.8 | 29.7 | 30.0 | 0.24 | 0.40 |
A3 | Zn0.5CeZrOx-glucose-600 | 20.5 | 20.4 | 20.9 | n.d. | 0.28 |
A4 | Zn0.5CeZrOx-glucose-700 | 20.2 | 20.2 | 20.0 | n.d. | 0.17 |
In addition, Fig. 3(III) shows the Ce 3d XPS spectra of the Zn0.5CeZrOx composite oxides. As Ce3+ was associated with the coordination-unsaturated O-substances on the ceria surface, the surface oxygen vacancy concentration could also be estimated using the ratio of Ce3+/(Ce3+ + Ce4+) from the surface concentration of Ce3+ and Ce4+ species determined from the Ce 3d XPS spectra.24–27 As given in Table 2, the surface O-defect concentration determined from the O 1s XPS spectra and the Ce3+ concentration determined from the Ce 3d XPS spectra display exactly the same trend as the surface oxygen vacancy concentration estimated from the Raman spectra. That is, the surface oxygen vacancies on the Zn0.5CeZrOx composite can be regulated by altering the complexing agent and calcination temperature; the Zn0.5CeZrOx-glucose-500 composite oxide, prepared with glucose as the complexing agent and calcined at 500 °C, has the most abundant surface oxygen vacancies.
During the sol–gel preparation process, various complexing agents may interact differently with the metal cations and hydrates of Zn, Ce and Zr, forming coordinate complexes different in structure and thereby composite oxides different in surface properties. It was reported that glucose as the complexing agent, owing to the abundant hydroxyl groups, could serve as the linkage to adsorb metal ions, forming nanospindles, fix these metal ions on the surface of glucose and thereby create abundant surface oxygen vacancies upon later calcination treatment,28,29 which may contribute to the high surface oxygen vacancy concentration of the Zn0.5CeZrOx-glucose composite oxide.
Moreover, the CO-TPD profiles shown in Fig. S9, ESI† also show that the Zn0.5CeZrOx-glucose composite oxide exhibits an intense desorption peak between 150 and 350 °C, in good accordance with the observation of Zhang, Zhou, Gao and co-workers that abundant surface oxygen vacancies could promote the adsorption and activation of CO on the composite oxides.30–32
When the Zn0.5CeZrOx composite oxides are combined with the SAPO-34 molecular sieves, as shown in Fig. 4(II), the bifunctional ZnCeZrO/SAPO-34 catalysts exhibit excellent performance in the conversion of syngas to olefins. Interestingly, the catalytic capacity of Zn0.5CeZrOx/SAPO-34 prepared with different complexing agents for producing light olefins from syngas follows the same trend as the capacity of Zn0.5CeZrOx for forming methanol (Fig. 4(III) and Table 2); that is, the bifunctional Zn0.5CeZrOx-glucose/SAPO-34 composite catalyst also gives a much higher selectivity to C2=–C4= (79.5%) and a larger space–time yield of C2=–C4= (0.4 mmol h−1 g−1) than the other composite catalysts.
Besides the complexing agent, the calcination temperature used in preparing the Zn0.5CeZrOx composite oxides also has a significant influence on the catalytic performance of Zn0.5CeZrOx/SAPO-34 in the STO reactions, as shown in Table 2 and Fig. S10, ESI.† In good accordance with the concentration of surface oxygen vacancies, the bifunctional Zn0.5CeZrOx-glucose-500/SAPO-34 catalyst, where Zn0.5CeZrOx-glucose-500 was prepared with glucose as the complexing agent and calcined at 500 °C, exhibits a much higher selectivity to C2=–C4= and a larger space–time yield of C2=–C4= than the other Zn0.5CeZrOx-glucose/SAPO-34 catalysts prepared by using a lower (400 °C) or a higher calcination temperature (600 and 700 °C).
The influence of the reaction temperature and ZnCeZrO/SAPO-34 mass ratio on the conversion of syngas to olefins over the bifunctional Zn0.5CeZrOx-glucose-500/SAPO-34 composite catalyst was also considered. As shown in Fig. 5(I), the selectivity to C2=–C4= is 64.1% at 280 °C, with a CO conversion of 5.9%, 7.0% for CH4, 4.5% for C20–C40 alkanes and 8.8% for CO2. With an increase in the reaction temperature, the conversion of CO increases gradually, but is accompanied with an increase in the selectivity to CO2 emission, due to the enhanced water-gas shift (WGS) reaction at higher temperature. The selectivity to C2=–C4= olefins presents a volcanic curve relationship with the reaction temperature and a maximum selectivity of 79.5% to C2=–C4= is achieved at 300 °C. Meanwhile, the ZnCeZrO/SAPO-34 mass ratio in the composite catalysts also has a significant influence on the product selectivity for STO, as shown in Fig. 5(II); under the current reaction conditions, the ZnCeZrO/SAPO-34 composite with an equal mass of Zn0.5CeZrOx and SAPO-34 is appropriate to give a high yield of light olefins for the conversion of syngas. A decrease in the content of the metal oxide moiety may lead to the decrease of CO conversion, whereas a decrease in the content of the molecular sieve moiety may not be able to realize complete evolution of the methanol intermediates into the olefin products.
Similarly, the reaction pressure, H2/CO ratio and space velocity of the syngas feed in the STO reaction were also optimized to maximize the formation of C2=–C4= olefins, as shown in Fig. S11, ESI.† It was demonstrated that for the STO reactions over the Zn0.5CeZrOx-glucose-500/SAPO-34 catalyst with an oxide/molecular sieve mass ratio of 1, a high space–time yield of C2=–C4= olefins can be achieved at 300 °C and 1.0 MPa, with a CO/H2 ratio of 1/2 and a syngas feed velocity of 5400 mL g−1 h−1. As shown in Fig. 5(III), for STO over the Zn0.5CeZrOx-glucose-500/SAPO-34 catalyst, the conversion of CO is kept at 6.5% during the 30 h reaction and the selectivities to CO2, CH4 and C20–C40 alkanes are lower than 10.7%, 5.5% and 4.1%, respectively, whereas the selectivity to C2=–C4= olefins increases gradually with the time on stream (TOS) from the initial 60.0% to 79.5% at a TOS of about 15 h and remains at this level in the later reaction stage, demonstrating the excellent performance of the bifunctional Zn0.5CeZrOx/SAPO-34 composite catalyst in the conversion of syngas to olefins.
After conducting the STO reactions, the crystal structure of the spent Zn0.5CeZrOx/SAPO-34 composite catalyst remains almost intact, as shown in Fig. S6, ESI,† although the surface area and pore volume decrease moderately in comparison with the fresh catalysts, due to the carbonaceous deposition on the acid sites of H-SAPO-34. Although SAPO-34 as a catalyst in MTO is somewhat notorious for its susceptibility to deactivation by the carbonaceous deposition, the Zn0.5CeZrOx/SAPO-34 composite catalyst exhibits relatively much better stability in STO (Fig. 5(III)). This is probably related to the low methanol formation rate over Zn0.5CeZrOx for syngas conversion (about 0.13–0.24 mmol h−1 g−1), which leads to a much lower concentration of methanol-related intermediate species on SAPO-34 in STO compared to that in a typical MTO process.33 Moreover, the presence of high concentration hydrogen in the feed is also beneficial for suppressing the formation of polycyclic aromatics and thereby elevating the catalytic lifetime of SAPO-34 zeolite in STO.34
However, when the space–time yields (STYs) of methanol over Zn0.5CeZrOx and those of C2=–C4= olefins over Zn0.5CeZrOx/SAPO-34 composites are associated with the concentration of surface oxygen vacancies on Zn0.5CeZrOx, as shown in Fig. 6(II), a good linear relationship is observed. That is, the surface oxygen vacancies play an important role in the catalytic conversion of syngas; an increase in the surface oxygen vacancy concentration of Zn0.5CeZrOx can almost proportionally improve the formation of methanol-related intermediates from the syngas over the Zn0.5CeZrOx moiety, promote the evolution of these intermediates into olefin products over the SAPO-34 moiety, and then enhance the overall capacity of the Zn0.5CeZrOx/SAPO-34 bifunctional catalysts in the conversion of syngas to light olefins, agreeing well with previous reports.30–32 The Zn0.5CeZrOx-glucose-500 composite oxide, prepared with glucose as the complexing agent and calcined at 500 °C, is provided with the most abundant surface oxygen vacancies; thereafter, the Zn0.5CeZrOx-glucose-500/SAPO-34 composite also exhibits excellent catalytic performance in the conversion of syngas to light olefins.
The promoting effect of surface oxygen vacancies on the formation of methanol intermediates over the ZnCeZrO composite oxide was further validated by in situ diffuse reflectance infrared Fourier transform (DRIFT) spectroscopy, as shown in Fig. 7. After the exposure of ZnCeZrO composite oxides to syngas, the CO adsorbed on ZnCeZrO can quickly interact with the active H* species, forming formate and methoxy intermediates through successive hydrogenations. The peaks in the DRIFT spectra centered at 1367, 1469 and 1596 cm−1 are attributed to the CO stretching and CH2 bending vibrations of surface formate species.35–37 The formate species are quickly generated after the reaction for 1 min and their peak intensity increases gradually with the reaction time. Another two peaks at 1051 and 1127 cm−1 representing the methoxy species are also clearly observed, accompanied by the characteristic peaks of C–H asymmetric and symmetric stretching vibrations of surface formate and methoxy species at 2962, 2850, 2924 and 2713 cm−1. Obviously, the peak intensity of formate and methoxy species on the Zn0.5CeZrOx-glucose composite is much stronger than that on Zn0.5CeZrOx-citric, as the former Zn0.5CeZrOx-glucose is provided with much more surface oxygen vacancies that can promote the conversion of syngas.
The results indicate that the surface oxygen vacancies on the ZnCeZrO composite oxide play an important role in the catalytic conversion of syngas, whose concentration can be finely tuned through altering the complexing agent and calcination temperature. An increase in the concentration of surface oxygen vacancies on Zn0.5CeZrOx can almost proportionally improve the formation of methanol-related intermediates from the syngas over the Zn0.5CeZrOx moiety, promote the evolution of these intermediates into the olefin products over the SAPO-34 moiety, and then enhance the overall capacity of the bifunctional Zn0.5CeZrOx/SAPO-34 composite catalyst in the conversion of syngas into light olefins.
The surface oxygen vacancy concentration of Zn0.5CeZrOx composite oxide can be markedly elevated by preparation through a sol–gel method with glucose as the complexing agent and calcination at 500 °C. When combined with the SAPO-34 molecular sieves, the bifunctional Zn0.5CeZrOx-glucose-500/SAPO-34 catalyst with abundant surface oxygen vacancies exhibits excellent performance in the synthesis of light olefins directly from syngas under mild reaction conditions (300 °C and 1.0 MPa), with a space–time yield of 0.40 mmol h−1 g−1 and a high selectivity of 79.5% to C2=–C4= olefins, whereas the selectivities to CH4, C20–C40 alkanes and CO2 are below 5.5%, 4.1% and 10.7%, respectively.
The current results depict a clear relationship between the concentration of surface oxygen vacancies on the ZnCeZrO moiety and the performance of the bifunctional ZnCeZrO/SAPO-34 catalyst in STO and provide an effective measure to elevate the abundance of surface oxygen vacancies, which should be beneficial to the design of efficient catalysts in the direct synthesis of light olefins from syngas.
Footnote |
† Electronic supplementary information (ESI) available: More characterization results for the Zn0.5CeZrOx composite oxides prepared with different complexing agents and calcined at different temperatures; optimization of the reaction conditions for the conversion of syngas to olefins over the Zn0.5CeZrOx–glucose-500/SAPO-34 bifunctional catalyst. See DOI: 10.1039/d0cy01759k |
This journal is © The Royal Society of Chemistry 2021 |