Next Article in Journal
Ultra-High Refractive Index Sensing Structure Based on a Metal-Insulator-Metal Waveguide-Coupled T-Shape Cavity with Metal Nanorod Defects
Next Article in Special Issue
Reductive and Coordinative Effects of Hydrazine in Structural Transformations of Copper Hydroxide Nanoparticles
Previous Article in Journal
The Synergistic Effect of Zinc Ferrite Nanoparticles Uniformly Deposited on Silver Nanowires for the Biofilm Inhibition of Candida albicans
Previous Article in Special Issue
Synthesis and Electrochemical Energy Storage Applications of Micro/Nanostructured Spherical Materials
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

A Facile Method for Preparing UiO-66 Encapsulated Ru Catalyst and its Application in Plasma-Assisted CO2 Methanation

College of Physical Science and Technology, Dalian University, Dalian 116622, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2019, 9(10), 1432; https://doi.org/10.3390/nano9101432
Submission received: 12 September 2019 / Revised: 29 September 2019 / Accepted: 4 October 2019 / Published: 10 October 2019
(This article belongs to the Special Issue Novel Nanomaterials for Applications in Energy and Catalysis)

Abstract

:
With increasing applications of metal-organic frameworks (MOFs) in the field of gas separation and catalysis, the preparation and performance research of encapsulating metal nanoparticles (NPs) into MOFs (M@MOF) have attracted extensive attention recently. Herein, an Ru@UiO-66 catalyst is prepared by a one-step method. Ru NPs are encapsulated in situ in the UiO-66 skeleton structure during the synthesis of UiO-66 metal-organic framework via a solvothermal method, and its catalytic activity for CO2 methanation with the synergy of cold plasma is studied. The crystallinity and structural integrity of UiO-66 is maintained after encapsulating Ru NPs according to the X-ray diffraction (XRD), Fourier transform infrared spectroscopy (FTIR), and scanning electron microscopy (SEM). As illustrated by X-ray photoelectron spectroscopy (XPS), high resolution transmission electron microscopy (HRTEM), and mapping analysis, the Ru species of the hydration ruthenium trichloride precursor are reduced to metallic Ru NPs without additional reducing processes during the synthesis of Ru@UiO-66, and the Ru NPs are uniformly distributed inside the Ru@UiO-66. Thermogravimetric analysis (TGA) and N2 sorption analysis show that the specific surface area and thermal stability of Ru@UiO-66 decrease slightly compared with that of UiO-66 and was ascribed to the encapsulation of Ru NPs in the UiO-66 skeleton. The results of plasma-assisted catalytic CO2 methanation indicate that Ru@UiO-66 exhibits excellent catalytic activity. CO2 conversion and CH4 selectivity over Ru@UiO-66 reached 72.2% and 95.4% under 13.0 W of discharge power and a 30 mL·min−1 gas flow rate ( V H 2 : V C O 2 = 4 : 1 ), respectively. Both values are significantly higher than pure UiO-66 with plasma and Ru/Al2O3 with plasma. The enhanced performance of Ru@UiO-66 is attributed to its unique framework structure and excellent dispersion of Ru NPs.

1. Introduction

Metal-organic frameworks (MOFs) are a series of porous crystal materials self-assembled by metal ions and organic ligands through coordination bonds [1,2,3]. The properties of specific surface area, porosity, and tunable functional structure make MOFs promising candidates in many applications such as gas storage and separation, drug delivery, chemical sensing, and catalysis [4,5,6,7,8]. As a typical Zr-based MOF, UiO-66 was first synthesized by Cavka et al. [9] and named after the University of Oslo. It has a perfect Zr6O4(OH)4 octahedral framework structure and exhibits exceptional thermal stability. In addition, UiO-66 has been widely utilized in the separation and conversion of CO2 [6,10,11,12,13].
The utility of UiO-66 as a support for metal catalysts (M/UiO-66, M=Cu, Au, Pd, Pt, Ru, etc.) has been extensively studied by researchers recently [14,15,16,17,18,19]. For example, Milet et al. [18] reported Pt/UiO-66 catalysts prepared by the double solvent method: impregnation of the UiO-66 support with the aqueous solution of H2PtCl6, and then reduction with NaBH4 solution. The Pt/UiO-66 exhibited excellent performance for CO2 methanation. The CO2 conversion and CH4 selectivity were as high as 50% and 36% at 350 °C with a CO2:H2 molar ratio of 1:5.2 and 1650 h−1 gas hourly space velocity (GHSV), respectively. Compared with the simple impregnation method, encapsulating metal nanoparticles (MNPs) into the skeleton structure of the MOF template can adjust the size distribution of MNPs and prepare high-performance MNPs@MOF catalysts [20]. Li et al. [21] synthesized UiO-66-encapsulated nano-palladium catalysts (Pd@UiO-66). As described, the Pd(acac)2 was first prepared, and then 1 g of UiO-66 and Pd(acac)2 solutions were mixed to synthesize Pd@UiO-66. Small Pd NPs (2.2 nm) were obtained due to the confinement of the small pore structure of UiO-66. The synthesized Pd@UiO-66 exhibited high catalytic activity and stability for continuous catalytic upgrading of ethanol to n-butanol. The ethanol conversion and the n-butanol selectivity over the best Pd@UiO-66 catalyst was 49.9% and 50.1%, respectively, during a 200-h evaluation. The high performance was attributed to the close synergy of highly distributed Pd NPs and coordinatively unsaturated Zr sites in UiO-66. Dong et al. [22] encapsulated Pd NPs in UiO-66 with a microwave-assisted method. The pores of UiO-66 were activated, and the metal precursors were reduced at the same time in the presence of NaBH4. The obtained Pd@UiO-66 exhibited high catalytic activity for Suzuki–Miyaura coupling reactions at mild conditions. Therefore, using MOFs with tunable porous structures as supports, growth of the metal NPs could be confined due to the encapsulation [23]. Consequently, high-performance MOFs supported metal catalysts with small size and high dispersion metal NPs can be generally obtained [24,25,26]. However, to synthesize the above MOFs-supported metal catalysts, there are two or more steps required, the synthesis of the MOF support and the impregnation, and the reduction of the supported/encapsulated metal precursors in the presence of reducing agents [18]. Therefore, this process is generally sophisticated and time-consuming.
The CO2 methanation has great prospects in economic and environmental applications since most of the fuel resources and one-carbon molecules (C1) can be regenerated from CO2 [27,28]. The emerging plasma-assisted activation of CO2 for methanation can provide the high energy for CO2 decomposition and overcome the relatively harsh conditions and reaction devices required for conventional thermochemical conversion [29,30,31,32,33]. Ru-based catalysts, due to their efficient activity, have been applied in CO2 methanation extensively. This is because, in the process of CO2 hydrogenation, the active Ru sites are more selective to the formation of methane, which can promote the reaction of CO2 along the methanation path in the hydrogen-rich environment [34,35,36,37].
In this work, we report a simple and efficient method for synthesizing Ru@UiO-66 via in situ encapsulation of Ru NPs by the reduction of the RuCl3 precursor during the growth of UiO-66 framework structure. The crystallinity and structural integrity of Ru@UiO-66 are similar to UiO-66. The synthesized Ru@UiO-66 exhibits high performance for plasma-assisted catalytic CO2 methanation.

2. Experimental

2.1. Materials

Zirconium tetrachloride (ZrCl4, 98%), terephthalic acid (BDC, 99%) and hydrochloric acid (HCl, 37%) were supplied by Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China). N,N-dimethylformamide (DMF, 99%) and hydration ruthenium trichloride (RuCl3·xH2O, Ru content 37%) were purchased from Tianjin Zhiyuan Co., Ltd. (Tianjin, China) and Walixi Chemical Co., Ltd. (Guangdong, China), respectively. Anhydrous methanol (CH3OH, 99.5%) was bought from Tianjin Kermel Co., Ltd. (Tianjin, China). All chemicals were used without further purification. High purity H2 (>99.999%) was generated by an HGH-500E hydrogen generator, and high purity CO2 and Ar (>99.999%) were obtained from the Guangming Research & Design Institute of Chemical Industry Co., Ltd. (Dalian, China).

2.2. Synthesis of UiO-66 and Ru@UiO-66

The strategy for preparing UiO-66 and Ru@UiO-66 is depicted in Figure 1. UiO-66 was prepared by a solvent thermal method previously reported in [38]. In brief, 1.165 g of ZrCl4 and 0.831 g of BDC were dissolved in 30 mL DMF, and then 0.8 mL of concentrated HCl (37%) was added. The obtained mixture was placed in an ultrasonic reactor for 20 min and heated at 120 °C for 24 h in a Teflon-lined steel autoclave. After cooling to room temperature, the obtained suspension of UiO-66 was centrifuged by a centrifuge machine (9000 rpm, 10 min), and rinsed with 25 mL of DMF and 25 mL of anhydrous methanol three times, respectively. The obtained UiO-66 was dried at 100 °C under vacuum. The synthesis procedure of Ru@UiO-66 was similar to that of UiO-66. The difference was that 0.415 g of RuCl3 with 1.165 g of ZrCl4 and 0.831 g of BDC were added into 30 mL of DMF. The other parameters were the same as that for synthesizing the pure UiO-66. The mass fraction of Ru determined by inductively coupled plasma optical emission spectrometer (ICP-OES) in Ru@UiO-66 was 2.83 wt%.

2.3. Catalytic Evaluation

The catalytic activity of the UiO-66 supported Ru catalysts was evaluated via a dielectric barrier discharge (DBD) plasma-assisted catalytic CO2 methanation reaction. The DBD plasma reactor is composed of a coaxial quartz tube (inner diameter: 8 mm, outer diameter: 10 mm), copper rod inner electrode (diameter: 2 mm) and 1 mm of a copper coil ground electrode. The discharge length and discharge gap are 25 mm and 2.5 mm, respectively. Typically, 0.3 g of samples were placed in the discharge area at a sinusoidal peak-to-peak voltage of 19.2 kV. The working gases of CO2 and H2 were mixed into the reactor after being measured by the mass flow meter (total flow rate was 30 mL·min−1, VH2:VCO2 = 4:1), while the outlet gas flow rate was measured by a soap film bubble flow meter. The gaseous products were analyzed online by a gas chromatograph (Tianmei GC-7890, Shanghai, China) equipped with a thermal conductivity detector (TCD). A type of FLUKE MT4 Max+ IR thermometer was used to monitor the temperature during the discharge process. The temperature of the reaction was ca. 200 °C. To evaluate the performance of CO2 methanation under cold plasma-assistance, CO2 conversion (XCO2), selectivity (S), and yield (Y) of the products were calculated as the reference [39]:
X C O 2 = F C O 2 F C O 2 F C O 2 × 100 %
S C O = F C O F C O 2 F C O 2 × 100 %
S C H 4 = F C H 4 F C O 2 F C O 2 × 100 %
Y = X × S × 100 %
where F and F′ are the inlet and outlet gas flow rates, respectively.

2.4. Catalysts Characterization

The X-ray diffraction (XRD) patterns were completed using a DX-2700 (Dandong, China) diffractometer with Cu Kα radiation (λ = 1.54178 Å) at 40 kV and 30 mA, and the step size of the measurement was 0.03°. Fourier transform infrared (FTIR) spectra were recorded in the range of 400–4000 cm−1 on a Nicolet AVATAR 370 (Waltham, MA, USA) infrared spectrometer. A Zeiss Sigma 500 (Jena, Germany) scanning electron microscope (SEM), operating at 10 kV, was utilized to characterize the morphology of the sample crystals. The X-ray photoelectron spectroscopy (XPS) spectra were recorded on a Thermo ESCALAN 250 spectrometer (Waltham, MA, USA) with Al Kα (1486.6 eV). The high-resolution transmission electron microscopy (HRTEM) and mapping were measured on a Tecnai G2 f20 s-twin (Hillsboro, OR, USA) transmission electron microscope. Binding energies were calibrated using C1s (284.6 eV) as the standard. The spectra were deconvoluted by the XPSPEAK41 program. N2 adsorption-desorption isotherms of the samples were performed on a NOVA 2200e analyzer (Boynton Beach, FL, USA) at a temperature of 77 K. The thermal stability of the samples was recorded by a Mettler TGA/DSC3+ thermal analyzer (Schwerzenbach, Switzerland) with a heating rate of 5 °C·min−1. Inductively coupled plasma optical emission spectrometer (ICP-OES) determination was executed on an Agilent 700 (Santa Clara, CA, USA) instrument.

3. Results and Discussion

Figure 2 illustrates the XRD patterns of Ru@UiO-66 and UiO-66. Obviously, the main diffraction peaks of Ru@UiO-66 at 2θ = 7.3°, 8.5° and 25.7° match well with UiO-66, revealing that no significant loss is observed in UiO-66 crystallinity after the introduction of RuCl3 in the synthesis of UiO-66. No diffraction peaks corresponding to Ru species can be detected in Ru@UiO-66, which may be ascribed to the low loading amount of Ru species, and/or the encapsulation of the Ru species into the UiO-66 skeleton structure.
FTIR spectra of Ru@UiO-66 and UiO-66 were measured, as illustrated in Figure 3. The FTIR spectra of Ru@UiO-66 and UiO-66 display the same peaks in the region of 4000 to 400 cm−1. The broad absorption band in the region of 3700 to 3200 cm−1 corresponding to the stretching vibration of O-H, to a large extent, is attributed to the residual solvent and adsorbed water [1,40]. In addition, it can also be induced by the OH from deprotonation of the carboxylate groups. The bands centered at ca. 1700 cm−1 and 1400 cm−1 correspond to the symmetrical stretching vibrations of the C=O bond in the -COO- group in the framework. The bands at 1506 cm−1 and 1581 cm−1 are assigned to the C=C stretching vibration of the phenyl ring. These indicate that the main functional groups in the BDC organic linker have been kept for Ru@UiO-66 and UiO-66. Furthermore, the peak at 745 cm−1 is consistent with the symmetric vibration peak of O-Zr-O and the symmetric vibration peak of O-Zr-O at 663 cm−1 [7,10]. All spectra have weak absorption bands in the region of 600–400 cm−1 pertaining to the in-plane and out-of-plane bending vibrations of -COO-. It was proven that Zr, as the coordination center in the organic framework, is formed by bridging with the organic ligand terephthalic acid through the bond of μ3-O [41]. However, the vibration peaks related to ruthenium cannot be detected from the FTIR spectra, suggesting that the ruthenium does not exist as a combined state in the UiO-66 skeleton.
The typical SEM images of Ru@UiO-66 and UiO-66 are shown in Figure 4. Ru@UiO-66 exhibits a similar morphology with UiO-66 in spite of the introduction of ruthenium in its skeleton. Furthermore, there are no obvious differences in the grain sizes of Ru@UiO-66 and UiO-66 according to the SEM images. The average grain size distribution for Ru@UiO-66 (202 ± 29 nm) is close to UiO-66 (201 ± 22 nm). Both values demonstrate that in situ addition of the Ru precursor during the fabrication of UiO-66 has no effect on the formation and uniform growth of the grains of Ru@UiO-66.
The XPS spectra of survey, Ru3p, Zr3d, and Cl2p over Ru@UiO-66 are shown in Figure 5. The elements of carbon, oxygen zirconium, and ruthenium are observed clearly in Figure 5a. Among these elements, carbon, oxygen, and zirconium constitute the catalysts. It is noted in Figure 5b that the peak appearing at 462.4 eV is attributed to 3p3/2 of Ru0 [36]. This reveals that the Ru3+ ions from RuCl3 have been reduced into metallic Ru0 directly, without extra reduction processes. During the synthesis of Ru@UiO-66, BDC and DMF serve as the skeleton of UiO-66 and solvent, respectively, and both of them can provide redundant H species for reducing Ru3+ ions. As validation, Ni@UiO-66, Pt@UiO-66, and Pd@UiO-66 have been successfully synthesized by the same method. The peak at 475 eV might be the Auger electron of O KL1L1 and KL1L23, which does not belong to any Ru species [42]. As mentioned in Figure 4, the peaks related to ruthenium are not detected in the FTIR spectrum of Ru@UiO-66, which further suggests that ruthenium existed in the framework of Ru@UiO-66 as a metallic state. The Zr3d5/2 and Zr3d3/2 in Ru@UiO-66 at 182.7 eV and 185.1 eV are attributed to the Zr4+ in O-Zr-O [7], which is also confirmed by the FTIR result (Figure 3). As shown in Figure 5c, no Cl can be detected, suggesting that the Cl- ions have been removed by washing during preparation of Ru@UiO-66.
The HRTEM and mapping measurements were further completed to observe the structure of Ru@UiO-66 as presented in Figure 6. Figure 6a,b clearly display that Ru NPs are regularly encapsulated in the framework of UiO-66 and {101} lattice fringes with an interplanar spacing of 0.204 nm [43]. This is in accordance with the low content of Ru0 on surface XPS spectra in Figure 5. The Ru NPs size distribution is presented in Figure 6c, it shows that Ru NPs have a small particle diameter of 1.66 nm and central distribution in the framework of UiO-66. Mapping analysis in random areas in Figure 6d–h reveals that C, O, and Zr elements abundantly exist in Ru@UiO-66, which are the main components of UiO-66 skeleton. In addition, Ru elements are evenly distributed in Ru@UiO-66.
Figure 7 gives the nitrogen adsorption-desorption isotherms and the classical Barrett–Joyner–Halenda (BJH) pore size distribution curves of Ru@UiO-66 and UiO-66. The specific surface area of Ru@UiO-66 is 766.4 m2·g−1, as calculated by the Brunauer–Emmett–Teller (BET) method, which is slightly lower than that of UiO-66 (996.9 m2·g−1). The pore diameter of Ru@UiO-66 (3.4 nm) is equal to UiO-66. This indicates that the addition of the precursor RuCl3 in the synthesis of Ru@UiO-66 did not affect the growth of the UiO-66 skeleton structure, and the pore size and pore volume of Ru@UiO-66 are similar to that of UiO-66.
To evaluate the thermal stability of Ru@UiO-66 and UiO-66, thermo-gravimetric analysis (TGA) was carried out in air, as shown in Figure 8. The weight loss of the three samples under 100 °C is attributed to desorption of water and residual solvent adsorbed at the surface. The TGA curve of UiO-66 declines rapidly at about 450 °C, suggesting the decomposition of the organic linker in the framework [8,44,45]. In contrast, the decomposition temperature of Ru@UiO-66 (ca. 350 °C) is lower than that of UiO-66 (450 °C), which may result from the presence of the Ru species. In other words, the interaction between -COO and Zr4+ is weakened due to the existence of the Ru species, and the Ru species facilitate the thermal decomposition of the framework at high temperature.
The activity of Ru@UiO-66 and UiO-66 for plasma-assisted catalytic CO2 methanation were tested, as illustrated in Figure 9. For comparison, the activity data of Ru/UiO-66 (UiO-66 supported Ru prepared by incipient wetness impregnation method) taken from a previous work [38] was also illustrated in Figure 9. The loading amount of Ru in Ru@UiO-66 and Ru/UiO-66 are close, and the reaction was performed under the same experimental parameters. The CO2 conversion of Ru@UiO-66 is 3.6 times higher than UiO-66, and the CH4 selectivity over Ru@UiO-66 slightly increases and reaches 95.4% at steady state. Even more interesting is that the products’ selectivities of the two samples are quite different. The primary product of Ru@UiO-66 is CH4, and the yield of CH4 reaches 68.9%. While the product of UiO-66 was mainly CO, the yield of CH4 was less than 3%. Due to the one-step preparation of Ru@UiO-66, the ruthenium in the precursor RuCl3 is directly reduced to the metallic Ru and encapsulated in the skeletal structure of the UiO-66. Therefore, the Ru@UiO-66 catalyst with active Ru species exhibits a perfect performance for plasma-assisted catalytic CO2 methanation. In addition, the CO2 conversion over Ru/UiO-66 increased from 20.4% for pure UiO-66 to 41.3%. However, it is still much lower than that over the Ru@UiO-66 catalyst (72.2%). Moreover, the CH4 selectivity over Ru/UiO-66 reaches 86.5%, while it is only ca. 3% for pure UiO-66. In summary, the Ru/UiO-66 possess much higher CO2 conversion activity, CH4 selectivity, and yield than pure UiO-66 due to the appearance of active Ru species. However, they exhibit poorer performance than the Ru@UiO-66 catalyst, which was mainly attributed to the encapsulation of highly dispersed active Ru species in the UiO-66 skeleton. These indicate that not only the Ru species but also the preparation methods play important roles in plasma-assisted CO2 conversion reactions.
The synergy of cold plasma and active Ru species play a vital role in the catalytic activity of CO2 methanation. Lee et al. [34] reported that the conversion of 5.369 wt% Ru/Al2O3 reached 23.2% under the synergy of DBD plasma. This is beyond the conversion of pure plasma and catalysts. Although this is the first report about the co-activation and conversion of CO2 by Ru-based catalysts with plasma, the conversion of CO2 is unsatisfactory. As for UiO-66, with a strong adsorption ability of CO2, it can serve as the carrier of the metal catalyst to effectively improve CO2 conversion. Unlike Ru/UiO-66, the active Ru NPs are distributed in the framework of Ru@UiO-66 uniformly to participate in the CO2 methanation effectively. The high catalytic activity of Ru@UiO-66 is due to the high dispersion of Ru NPs and the high CO2 adsorption ability of the UiO-66 skeleton.

4. Conclusions

In summary, Ru@UiO-66 was successfully synthesized via a one-step solvothermal method. The Ru species from the RuCl3 precursor was reduced directly and embedded into the framework of UiO-66 during the synthesis of Ru@UiO-66, without an additional reducing process. The crystallinity and structural integrity of UiO-66 were maintained well after encapsulating Ru NPs, and the Ru NPs were uniformly distributed inside the framework of UiO-66. The results of the plasma-assisted CO2 methanation indicated that Ru@UiO-66 exhibited an unprecedented catalytic activity with the synergy of plasma. CO2 conversion and CH4 selectivity over Ru@UiO-66 reached 72.2% and 95.4% under 13.0 W discharge power and 30 mL·min−1 gas flow rate ( V H 2 : V C O 2 = 4 : 1 ). The high performance of Ru@UiO-66 can be ascribed to the synergy of the Ru NPs and cold plasma. The high CO2 adsorption ability of UiO-66 is essential for CO2 methanation. This work provides a simple method to synthesize high-performance MOF-supported Ru catalysts via a one-step solvothermal method to reduce and encapsulate the Ru NPs into the skeleton at the same time.

Author Contributions

Conceptualization, L.D. and X.Z.; Data curation, W.X. and M.D.; Investigation, W.X.; Project administration, L.D.; Supervision, L.D. and X.Z.; Writing—original draft, W.X.; Writing—review & editing, X.Z.

Funding

This work is supported by National Natural Science Foundation of China (Nos. 21673026, 21773020, 11505019), the Liaoning Innovative Talents in University (No. LR2017025), the Liaoning Natural Science Foundation (No. 20180550085), and Graduate Education and Teaching Reform Fund of Dalian University.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yang, F.; Li, W.; Tang, B. Facile synthesis of amorphous UiO-66 (Zr-MOF) for supercapacitor application. J. Alloy. Compd. 2018, 733, 8–14. [Google Scholar] [CrossRef]
  2. Rungtaweevoranit, B.; Baek, J.; Araujo, J.R.; Archanjo, B.S.; Choi, K.M.; Yaghi, O.M.; Somotjai, G.A. Copper nanocrystals encapsulated in Zr-based metal-organic frameworks for highly selective CO2 hydrogenation to methanol. Nano Lett. 2016, 16, 7645–7649. [Google Scholar] [CrossRef] [PubMed]
  3. Li, R.; Zhang, W.; Zhou, K. Metal-organic-framework-based catalysts for photoreduction of CO2. Adv. Mater. 2018, 30, 1705512. [Google Scholar] [CrossRef] [PubMed]
  4. Li, B.; Wen, H.-M.; Zhou, W.; Chen, B. Porous metal–organic frameworks for gas storage and separation: What, How, and Why? J. Phys. Chem. Lett. 2014, 5, 3468–3479. [Google Scholar] [CrossRef] [PubMed]
  5. Abid, H.R.; Ang, H.M.; Wang, S. Effects of ammonium hydroxide on the structure and gas adsorption of nanosized Zr-MOFs (UiO-66). Nanoscale 2012, 4, 3089–3094. [Google Scholar] [CrossRef]
  6. Wang, Y.; Hu, Z.; Kundu, T.; Cheng, Y.; Dong, J.; Qian, Y.; Zhai, L.; Zhao, D. Metal-organic frameworks with reduced hydrophilicity for postcombustion CO2 capture from wet flue gas. ACS Sustain. Chem. Eng. 2018, 6, 11904–11912. [Google Scholar] [CrossRef]
  7. Jiang, X.; Li, S.; He, S.; Bai, Y.; Shao, L. Interface manipulation of CO2-philic composite membranes containing designed UiO-66 derivatives towards highly efficient CO2 capture. J. Mater. Chem. 2018, 6, 15064–15073. [Google Scholar] [CrossRef]
  8. Denny, M.S., Jr.; Parent, L.R.; Patterson, J.P.; Meena, S.K.; Pham, H.; Abellan, P.; Ramasse, Q.M.; Paesani, F.; Gianneschi, N.C.; Cohen, S.M. Transmission electron microscopy reveals deposition of metal oxide coatings onto metal-organic frameworks. J. Am. Chem. Soc. 2018, 140, 1348–1357. [Google Scholar] [CrossRef]
  9. Cavka, J.H.; Jakobsen, S.; Olsbye, U.; Guillou, N.; Lamberti, C.; Bordiga, S.; Lillerud, K.P. A new zirconium inorganic building brick forming metal organic frameworks with exceptional stability. J. Am. Chem. Soc. 2008, 130, 13850–13851. [Google Scholar] [CrossRef]
  10. Rodrigues, M.A.; de Souza Ribeiro, J.; de Souza Costa, E.; de Miranda, J.L.; Ferraz, H.C. Nanostructured membranes containing UiO-66 (Zr) and MIL-101 (Cr) for O2/N2 and CO2/N2 separation. Sep. Purif. Technol. 2018, 192, 491–500. [Google Scholar] [CrossRef]
  11. Hu, Z.; Khurana, M.; Seah, Y.H.; Zhang, M.; Guo, Z.; Zhao, D. Ionized Zr-MOFs for highly efficient post-combustion CO2 capture. Chem. Eng. Sci. 2015, 124, 61–69. [Google Scholar] [CrossRef]
  12. Huang, H.; Zhang, W.; Yang, F.; Wang, B.; Yang, Q.; Xie, Y.; Zhong, C.; Li, J.-R. Enhancing CO2 adsorption and separation ability of Zr (IV)-based metal-organic frameworks through ligand functionalization under the guidance of the quantitative structure-property relationship model. Chem. Eng. J. 2016, 289, 247–253. [Google Scholar] [CrossRef]
  13. Hu, J.; Liu, Y.; Liu, J.; Gu, C.; Wu, D. High CO2 adsorption capacities in UiO type MOFs comprising heterocyclic ligand. Microporous Mesoporous Mater. 2018, 256, 25–31. [Google Scholar] [CrossRef]
  14. Abdel-Mageed, A.M.; Rungtaweevoranit, B.; Parlinska-Wojtan, M.; Pei, X.; Yaghi, O.M.; Behm, R.J. Highly active and stable single-atom Cu catalysts supported by a metal-organic framework. J. Am. Chem. Soc. 2019, 141, 5201–5210. [Google Scholar] [CrossRef] [PubMed]
  15. Qiu, J.; Zhang, X.; Xie, K.; Zhang, X.-F.; Feng, Y.; Jia, M.; Yao, J. Noble metal nanoparticle-functionalized Zr-metal organic frameworks with excellent photocatalytic performance. J. Colloid Interface Sci. 2019, 538, 569–577. [Google Scholar] [CrossRef]
  16. Limvorapitux, R.; Chou, L.-Y.; Young, A.P.; Tsung, C.-K.; Nguyen, S.T. Coupling molecular and nanoparticle catalysts on single metal-organic framework microcrystals for the tandem reaction of H2O2 generation and selective alkene oxidation. ACS Catal. 2017, 7, 6691–6698. [Google Scholar] [CrossRef]
  17. Bakuru, V.R.; Velaga, B.; Peela, N.R.; Kalidindi, S.B. Hybridization of Pd nanoparticles with UiO-66 (Hf) metal-organic framework and the effect of nanostructure on the catalytic properties. Chem. Eur. J. 2018, 24, 15978–15982. [Google Scholar] [CrossRef]
  18. Mihet, M.; Blanita, G.; Dan, M.; Barbu-Tudoran, L.; Lazar, M.D. Pt/UiO-66 nanocomposites as catalysts for CO2 methanation process. J. Nanosci. Nanotechnol. 2019, 19, 3187–3196. [Google Scholar] [CrossRef]
  19. Li, Z.; Rayder, T.M.; Luo, L.; Byers, J.A.; Tsung, C.K. Aperture-opening encapsulation of a transition metal catalyst in a metal-organic framework for CO2 hydrogenation. J. Am. Chem. Soc. 2018, 140, 8082–8085. [Google Scholar] [CrossRef]
  20. Deng, M.; Bo, X.; Guo, L. Encapsulation of platinum nanoparticles into a series of zirconium-based metal-organic frameworks: Effect of the carrier structures on electrocatalytic performances of composites. J. Electroanal. Chem. 2018, 815, 198–209. [Google Scholar] [CrossRef]
  21. Jiang, D.; Fang, G.; Tong, Y.; Wu, X.; Wang, Y.; Hong, D.; Leng, W.; Liang, Z.; Tu, P.; Liu, L.; et al. Multifunctional Pd@UiO-66 catalysts for continuous catalytic upgrading of ethanol to n-butanol. ACS Catal. 2018, 8, 11973–11978. [Google Scholar] [CrossRef]
  22. Dong, W.; Feng, C.; Zhang, L.; Shang, N.; Gao, S.; Wang, C.; Wang, Z. Pd@UiO-66: An efficient catalyst for Suzuki-Miyaura coupling reaction at mild condition. Catal. Lett. 2016, 146, 117–125. [Google Scholar] [CrossRef]
  23. Luz, I.; Roesler, C.; Epp, K.; Llabres i Xamena, F.X.; Fischer, R.A. Pd@UiO-66-type MOFs prepared by chemical vapor infiltration as shape-selective hydrogenation catalysts. Eur. J. Inorg. Chem. 2015, 3904–3912. [Google Scholar] [CrossRef]
  24. Xiao, J.-D.; Shang, Q.; Xiong, Y.; Zhang, Q.; Luo, Y.; Yu, S.-H.; Jiang, H.-L. Boosting photocatalytic hydrogen production of a metal-organic framework decorated with platinum nanoparticles: The platinum location matters. Angew. Chem. Int. Ed. 2016, 55, 9389–9393. [Google Scholar] [CrossRef] [PubMed]
  25. Ling, P.; Lei, J.; Jia, L.; Ju, H. Platinum nanoparticles encapsulated metal-organic frameworks for the electrochemical detection of telomerase activity. Chem. Commun. 2016, 52, 1226–1229. [Google Scholar] [CrossRef] [PubMed]
  26. Li, B.; Ma, J.-G.; Cheng, P. Integration of metal nanoparticles into metal–organic frameworks for composite catalysts: Design and synthetic strategy. Small 2019, 1804849. [Google Scholar] [CrossRef] [PubMed]
  27. Jia, X.; Zhang, X.; Rui, N.; Hu, X.; Liu, C.-J. Structural effect of Ni/ZrO2 catalyst on CO2 methanation with enhanced activity. Appl. Catal. B Environ. 2019, 244, 159–169. [Google Scholar] [CrossRef]
  28. Mei, D.; Tu, X. Conversion of CO2 in a cylindrical dielectric barrier discharge reactor: Effects of plasma processing parameters and reactor design. J. CO2 Util. 2017, 19, 68–78. [Google Scholar] [CrossRef]
  29. Alliati, M.; Mei, D.; Tu, X. Plasma activation of CO2 in a dielectric barrier discharge: A chemical kinetic model from the microdischarge to the reactor scales. J. CO2 Util. 2018, 27, 308–319. [Google Scholar] [CrossRef]
  30. Sun, K.; Fan, Z.; Ye, J.; Yan, J.; Ge, Q.; Li, Y.; He, W.; Yang, W.; Liu, C.-J. Hydrogenation of CO2 to methanol over In2O3 catalyst. J. CO2 Util. 2015, 12, 1–6. [Google Scholar] [CrossRef]
  31. Fan, Z.; Sun, K.; Rui, N.; Zhao, B.; Liu, C.-J. Improved activity of Ni/MgAl2O4 for CO2 methanation by the plasma decomposition. J. Energy Chem. 2015, 24, 655–659. [Google Scholar] [CrossRef]
  32. Zhou, A.; Chen, D.; Dai, B.; Ma, C.; Li, P.; Yu, F. Direct decomposition of CO2 using self-cooling dielectric barrier discharge plasma. Greenh. Gases Sci. Technol. 2017, 7, 721–730. [Google Scholar] [CrossRef]
  33. Zhou, A.; Chen, D.; Ma, C.; Yu, F.; Dai, B. DBD Plasma-ZrO2 catalytic decomposition of CO2 at low temperatures. Catalysts 2018, 8, 256. [Google Scholar] [CrossRef]
  34. Lee, C.J.; Lee, D.H.; Kim, T. Enhancement of methanation of carbon dioxide using dielectric barrier discharge on a ruthenium catalyst at atmospheric conditions. Catal. Today 2017, 293, 97–104. [Google Scholar] [CrossRef]
  35. Wu, T.B.; Zhang, P.; Ma, J.; Fan, H.L.; Wang, W.T.; Jiang, T.; Han, B.X. Catalytic activity of immobilized Ru nanoparticles in a porous metal-organic framework using supercritical fluid. Chin. J. Catal. 2013, 34, 167–175. [Google Scholar] [CrossRef]
  36. Wu, C.; Irshad, F.; Luo, M.; Zhao, Y.; Ma, X.; Wang, S. Ruthenium complexes immobilized on an azolium based metal organic framework for highly efficient conversion of CO2 into formic acid. ChemCatChem 2019, 11, 1256–1263. [Google Scholar] [CrossRef]
  37. Shang, X.; Deng, D.; Wang, X.; Xuan, W.; Zou, X.; Ding, W.; Lu, X. Enhanced low-temperature activity for CO2 methanation over Ru doped the Ni/CexZr(1-x) O2 catalysts prepared by one-pot hydrolysis method. Int. J. Hydrog. Energy 2018, 43, 7179–7189. [Google Scholar] [CrossRef]
  38. Xu, W.; Zhang, X.; Dong, M.; Zhao, J.; Di, L. Plasma-assisted Ru/Zr-MOF catalyst for hydrogenation of CO2 to methane. Plasma Sci. Technol. 2019, 21, 044004. [Google Scholar] [CrossRef]
  39. Wang, L.; Yi, Y.H.; Guo, H.C.; Tu, X. Atmospheric pressure and room temperature synthesis of methanol through plasma-catalytic hydrogenation of CO2. ACS Catal. 2018, 8, 90–100. [Google Scholar] [CrossRef]
  40. Mocniak, K.A.; Kubajewska, I.; Spillane, D.E.M.; Williams, G.R.; Morris, R.E. Incorporation of cisplatin into the metal-organic frameworks UiO66-NH2 and UiO66-encapsulation vs. conjugation. RSC Adv. 2015, 5, 83648–83656. [Google Scholar] [CrossRef]
  41. Pourkhosravani, M.; Dehghanpour, S.; Farzaneh, F. Palladium nanoparticles supported on zirconium metal organic framework as an efficient heterogeneous catalyst for the Suzuki-Miyaura coupling reaction. Catal. Lett. 2016, 146, 499–508. [Google Scholar] [CrossRef]
  42. Xu, G.; Zhang, Y.; Fu, Y.; Guo, Q. Efficient hydrogenation of various renewable oils over Ru-HAP catalyst in water. ACS Catal. 2017, 7, 1158–1169. [Google Scholar] [CrossRef]
  43. Wang, F.; Li, C.; Zhang, X.; Wei, M.; Evans, D.G.; Duan, X. Catalytic behavior of supported Ru nanoparticles on the {100}, {110}, and {111} facet of CeO2. J. Catal. 2015, 329, 177–186. [Google Scholar] [CrossRef]
  44. Jiang, H.; Gao, Q.; Wang, S.; Chen, Y.; Zhang, M. The synergistic effect of Pd NPs and UiO-66 for enhanced activity of carbon dioxide methanation. J. CO2 Util. 2019, 31, 167–172. [Google Scholar] [CrossRef]
  45. Zhang, S.; Li, L.; Zhao, S.; Sun, Z.; Luo, J. Construction of interpenetrated ruthenium metal-organic frameworks as stable photocatalysts for CO2 reduction. Lnorg. Chem. 2015, 54, 8375–8379. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic illustration for the synthesis of UiO-66 and Ru@UiO-66.
Figure 1. Schematic illustration for the synthesis of UiO-66 and Ru@UiO-66.
Nanomaterials 09 01432 g001
Figure 2. X-ray diffraction (XRD patterns of Ru@UiO-66 and UiO-66.
Figure 2. X-ray diffraction (XRD patterns of Ru@UiO-66 and UiO-66.
Nanomaterials 09 01432 g002
Figure 3. Fourier transform infrared spectroscopy (FTIR) spectra of Ru@UiO-66 and UiO-66.
Figure 3. Fourier transform infrared spectroscopy (FTIR) spectra of Ru@UiO-66 and UiO-66.
Nanomaterials 09 01432 g003
Figure 4. Scanning electron microscopy (SEM) images and the corresponding histograms of the size distribution of (a) Ru@UiO-66 and (b) UiO-66.
Figure 4. Scanning electron microscopy (SEM) images and the corresponding histograms of the size distribution of (a) Ru@UiO-66 and (b) UiO-66.
Nanomaterials 09 01432 g004
Figure 5. The X-ray photoelectron spectroscopy XPS spectra of (a) survey, (b) Ru3p, (c) Zr3d, and (d) Cl2p over Ru@UiO-66 and UiO-66.
Figure 5. The X-ray photoelectron spectroscopy XPS spectra of (a) survey, (b) Ru3p, (c) Zr3d, and (d) Cl2p over Ru@UiO-66 and UiO-66.
Nanomaterials 09 01432 g005
Figure 6. (a,b) The HRTEM images of Ru@UiO-66. (c) The size distribution histogram of Ru NPs. (dh) The mapping of C, O, Zr, Ru elements on Ru@UiO-66 catalyst.
Figure 6. (a,b) The HRTEM images of Ru@UiO-66. (c) The size distribution histogram of Ru NPs. (dh) The mapping of C, O, Zr, Ru elements on Ru@UiO-66 catalyst.
Nanomaterials 09 01432 g006
Figure 7. (a) Nitrogen adsorption-desorption isotherms and (b) the BJH pore size distribution curves of Ru@UiO-66 and UiO-66.
Figure 7. (a) Nitrogen adsorption-desorption isotherms and (b) the BJH pore size distribution curves of Ru@UiO-66 and UiO-66.
Nanomaterials 09 01432 g007
Figure 8. Thermogravimetric analysis (TGA) curves of Ru@UiO-66 and UiO-66.
Figure 8. Thermogravimetric analysis (TGA) curves of Ru@UiO-66 and UiO-66.
Nanomaterials 09 01432 g008
Figure 9. Dielectric barrier discharge (DBD) plasma-assisted CO2 conversion over UiO-66, Ru@UiO-66, and Ru/UiO-66, Reproduced from [38], Copyright Hefei Institutes of Physical Science, Chinese Academy of Sciences and IOP Publishing, 2019.
Figure 9. Dielectric barrier discharge (DBD) plasma-assisted CO2 conversion over UiO-66, Ru@UiO-66, and Ru/UiO-66, Reproduced from [38], Copyright Hefei Institutes of Physical Science, Chinese Academy of Sciences and IOP Publishing, 2019.
Nanomaterials 09 01432 g009

Share and Cite

MDPI and ACS Style

Xu, W.; Dong, M.; Di, L.; Zhang, X. A Facile Method for Preparing UiO-66 Encapsulated Ru Catalyst and its Application in Plasma-Assisted CO2 Methanation. Nanomaterials 2019, 9, 1432. https://doi.org/10.3390/nano9101432

AMA Style

Xu W, Dong M, Di L, Zhang X. A Facile Method for Preparing UiO-66 Encapsulated Ru Catalyst and its Application in Plasma-Assisted CO2 Methanation. Nanomaterials. 2019; 9(10):1432. https://doi.org/10.3390/nano9101432

Chicago/Turabian Style

Xu, Weiwei, Mengyue Dong, Lanbo Di, and Xiuling Zhang. 2019. "A Facile Method for Preparing UiO-66 Encapsulated Ru Catalyst and its Application in Plasma-Assisted CO2 Methanation" Nanomaterials 9, no. 10: 1432. https://doi.org/10.3390/nano9101432

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