Next Article in Journal
The Effect of Si on CO2 Methanation over Ni-xSi/ZrO2 Catalysts at Low Temperature
Previous Article in Journal
Special Issue “Selected Papers from the 5nd Edition of Global Conference on Catalysis, Chemical Engineering and Technology (CAT 2019)”
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ruthenium Nanoparticles Intercalated in Montmorillonite (nano-Ru@MMT) Is Highly Efficient Catalyst for the Selective Hydrogenation of 2-Furaldehyde in Benign Aqueous Medium

1
H.E.J. Research Institute of Chemistry, International Center for Chemical and Biological Sciences, University of Karachi, Karachi 75270, Pakistan
2
Center of Research Excellence in Nanotechnology, King Fahd University of Petroleum and Minerals (KFUPM), Dhahran 31261, Saudi Arabia
3
Department of Chemistry, King Fahd University of Petroleum and Minerals, Dhahran 31261, Saudi Arabia
4
Department of Chemistry, GC University Lahore, Lahore 54000, Pakistan
5
School of Chemistry, Chemical Engineering and Life Sciences, Wuhan University of Technology, Wuhan 430070, China
6
Chemistry Department, College of Science, Taif University, P.O. Box 11099, Taif 21944, Saudi Arabia
7
On Leave from Chemistry Department, Faculty of Science, Aswan University, P.O. Box 81528, Aswan 81528, Egypt
8
International Research Center for Food Nutrition and Safety, Jiangsu University, Zhenjiang 212013, China
9
Department of Molecular Biosciences, The Wenner-Gren Institute, Stockholm University, S-106 91 Stockholm, Sweden
10
Department of Chemistry, Faculty of Science, Menoufia University, Menoufia 32511, Egypt
*
Authors to whom correspondence should be addressed.
Catalysts 2021, 11(1), 66; https://doi.org/10.3390/catal11010066
Submission received: 1 December 2020 / Revised: 25 December 2020 / Accepted: 1 January 2021 / Published: 5 January 2021

Abstract

:
Chemoselective hydrogenation of 2-furaldehyde to furfuryl alcohol using green solvents is an important research area to get eco-friendly fuels and fine chemicals. Herein, we report ruthenium nanoparticles (~1.8 nm) intercalated in montmorillonite as an efficient catalytic system, which can selectively hydrogenate 2-furaldehyde in a benign aqueous medium. The complete conversion was observed at 40 °C with 1 MPa H2, the selectivity of furfuryl alcohol being >99%, and turnover number 1165. After a catalytic run, the montmorillonite-supported ruthenium nanoparticles can be recycled and reused without losing their activity and selectivity.

Graphical Abstract

1. Introduction

Hydrogenation of biomass-derived chemicals is considered an important route towards renewable energy resources [1,2]. Among bio-mass derived organics, 2-furaldehyde is a versatile source of many fine chemicals. Its hydrogenated product—furfuryl alcohol—is a key synthetic intermediate and feedstock for the manufacture of lysine, ascorbic acid, lubricants, corrosion-resistant fiberglass, thermostatic resins, acid proof bricks, and it is also being used in the polymer industry [1,3,4,5]. Thus, selective hydrogenation of 2-furaldehyde to furfuryl alcohol using environment-friendly solvents, such as water, is a fascinating area of research to fulfill the increasing demand for eco-friendly fuels and fine chemicals [6]. Using water as a solvent is also more consistent with “Green Chemistry” objectives. It is therefore important to develop a catalyst capable of converting 2-furaldehyde to furfuryl alcohol under mild conditions in a benign aqueous medium. A brief literature survey reveals that metallic ruthenium nanoparticles have the potential to selectively hydrogenate aldehydes and ketones into their respective alcohols [7]. Yuan et al. showed that 2-furaldehyde can be converted into furfuryl alcohol using the ruthenium nanoparticles (conversion 100%, selectivity 95%) [8]. Similarly, Chen et al. achieved this hydrogenation by employing ruthenium nanoparticles under mild conditions (water, 0.5 MPa H2, 25 °C), although catalyst selectivity was moderate [9]. Graphene oxide supported ruthenium nanoparticles (Ru/rGO) were shown to yield 91% furfuryl alcohol in water (1 MPa H2, 20 °C) [10]. Likewise, Yuan et al. found 94.9% furfuryl alcohol selectivity over ruthenium nanoparticles supported on a zirconium-based metal-organic framework (water, 0.5 MPa H2, 20 °C) [8]. However, ruthenium nanoparticles supported on a metal-organic framework having an aromatic linker (Al-MIL-53-BDC) could completely convert 2-furaldehyde to furfuryl alcohol under these conditions [11]. Similarly, G. Bagnato et al. reported a novel ruthenium-polyethersulfone type catalytic membrane (Ru/Ph2P(CH2)4PPh2), which could selectively produce furfuryl alcohol at 140 °C [12]. To sum up the ruthenium-based catalysts, neutral supports seem preferable for the production of furfuryl alcohol under mild reaction conditions in water. Kyriakou, G. and co-workers developed a platinum-based catalyst involving different oxide supports, such as MgO, Al2O3, ZnO, SiO2, and CeO2. Furfural hydrogenation was found to be sensitive toward platinum nanoparticle size and solvent choice. Their catalyst showed high activity and selectivity in alcohols. Non-polar solvents offered poor conversions and favored many side products [13]. Trimetallic Cu-Fe-Al-containing catalyst had an overall furfuryl alcohol yield of 97% in isopropyl alcohol at 100 °C and 6.0 MPa hydrogen pressure [14]. Gómez et al. prepared copper containing a heterogeneous catalyst using clay minerals (bentonite and sepiolite). 2-Furaladehyde conversion up to 83% was observed for Cu-Bentonite catalyst [15]. Overall, a number of catalysts have been reported in the literature for selective hydrogenation of 2-furaldehyde including Ni, Pt, Pd, Cu, Ru, Pd-Au, Pt-Sn, Cu-Ca, Cu-Fe-Al on various metallic, bimetallic, and amorphous supports [1,3,4,5,16,17,18]. However, commercial production of furfuryl alcohol is mainly achieved with copper chromate catalysts. Nevertheless, drastic temperature conditions (>100 °C) are employed, which results in the formation of bi-products mainly due to hydrogenolysis [19]. Although copper chromate exhibits high selectivity and activity, its application is limited due to high Cr(VI) toxicity and serious environmental concerns [3]. Thus, it is highly desirable to develop an environment-friendly catalyst with high activity and selectivity.
In heterogeneous catalysis, montmorillonite (MMT) is one of the most intensively explored catalytic materials [20]. This clay, as catalyst support, shows great potential in the design of green hydrogenation catalyst. Montmorillonite is dioctahedral smectite with an ideal unit cell formula of Na0.4(Al1.6Mg0.4)Si4O10(OH)2. It’s 2:1 phyllosilicate, in which an octahedral sheet of alumina is sandwiched between two tetrahedral silicate layers [21]. Interlaminar space usually contains exchangeable sodium cations (see Figure 1). Water-soluble organometallic cations, such as arene ruthenium complexes can easily intercalate the interlaminar space of smectite clays by replacing these sodium cations in an aqueous medium.
For hydrogenation reactions, ruthenium is considered a promising catalytic material. However, designing a selective catalyst necessitates a suitable metal precursor, which plays a crucial role in size, morphology, and individual molecular behavior on the surface of nanoparticles [23]. Nanoparticles with controllable chemical and physical properties can conveniently be obtained by employing suitable organometallic precursors [24]. We have previously shown that the selectivity during catalytic hydrogenation reactions can be conveniently controlled with nano ruthenium, which was obtained from arene ruthenium complexes devoid of interfering chloride ions [21,25,26,27,28,29,30,31]. In this work, dicationic aqua complex [Me(C6H4)Pri)Ru(H2O)3]2+ was used as a precursor to avoid the presence of contaminants on the surface of nanoparticles. Herein, we report a highly efficient catalyst (conversion 100%, selectivity >99%, turnover number (TON) 1165) containing montmorillonite-supported ruthenium nanoparticles (~1.8 nm), which can hydrogenate 2-furaldehyde into furfuryl alcohol under mild conditions (1 MPa H2, 40 °C) in a benign aqueous medium.

2. Results and Discussion

When dimeric ruthenium complex [Me(C6H4)Pri)2Ru2Cl4] is dissolved in water, the solution contains three complexes [Me(C6H4)Pri)RuCl2(H2O)], [Me(C6H4)Pri)Ru(H2O)3]2+, and [Me(C6H4)Pri)RuCl(H2O)2]+ in equilibrium (Scheme 1). However, this equilibrium can be shifted towards dicationic aqua complex [Me(C6H4)Pri)Ru(H2O)3]2+ over a pH range of 5–8, just like in the case of dimeric benzene ruthenium complex [(C6H6)2Ru2Cl4] [32].
When the yellow aqueous solution of [Me(C6H4)Pri)2Ru2Cl4], predominantly containing [Me(C6H4)Pri)Ru(H2O)3]2+ after pH adjustment to 8, is added to sodium montmorillonite (1), the dicationic aqua complex [Me(C6H4)Pri)Ru(H2O)3]2+ intercalates the lamellae of 1 by ion-exchange with sodium cations giving the yellow Ru(II)-modified montmorillonite 2. Reduction of the dicationic aqua complex in 2 with hydrogen (10 bar) at 100 °C in ethanol affords black Ru(0)-modified montmorillonite 3 (Scheme 2).
BET surface area of 3, determined by low-temperature nitrogen adsorption, was 38.60 m2g−1 whereas the average pore diameter was 75 Å (Figure 2). Ruthenium loading in 3 (3.0 wt.%) was calculated from the molar ratio of [Me(C6H4)Pri)Ru(H2O)3]2+ used, which corresponds to 75 percent CEC of 1. Ruthenium loading determined by ICP-MS was found to be in agreement with the calculated value. A transmission electron microscope (TEM) was used to determine the overall size distribution of nanoparticles in Ru(0)-modified montmorillonite 3. To calculate the particle size in TEM micrographs, 250 particles were chosen. Selected area electron diffraction (SAED) ascertained the crystalline nature of these nanoparticles. TEM micrographs were processed with “ImageJ” software, which revealed the nanoparticles having an average size of 1.8 nm with a standard deviation (σ) being less than 25%. Scanning electron microscope (SEM) was used to study the catalyst morphology and SEM coupled energy-dispersive X-ray spectroscopy (EDXS) confirmed the presence of ruthenium in 3. X-ray powder diffraction (XRD) pattern comparison of Ru(0)-containing montmorillonite 3 with that of unmodified montmorillonite 1 is shown in Figure 2, the high basal spacing value (d001 = 17.1 Å) of 3 can be related to the expansion of interlaminar space due to ethanol.
Montmorillonite intercalated ruthenium nanoparticles 3 demonstrate high selectivity and high activity for the hydrogenation of 2-furaldehyde. However, catalyst activity is highly dependent on the way Ru(0)-montmorillonite 3 is prepared. When the catalyst is prepared by reducing Ru(II)-montmorillonite 2 in water (1 MPa H2, 100 °C), the resulting catalyst 3 shows poor activity. TEM micrographs revealed the presence of ruthenium nanoparticles with an average size of 12 nm (see Figure 3). Conversely, the reduction of 2 in ethanol (1 MPa H2, 100 °C) affords 3, which demonstrates high activity and selectivity for the hydrogenation of 2-furaldehyde in water. This catalyst affords furfuryl alcohol selectively under suitable mild conditions while avoiding the formation of side products. Now TEM analysis indicates the existence of smaller-sized (1.8 nm) Ru(0)-nanoparticles in montmorillonite, which leads us to the conclusion that 2-furaldehyde conversion and furfuryl alcohol selectivity is mainly dependent on the size of ruthenium nanoparticles. Moreover, previous literature suggests that an aromatic π cloud can act as an H-bond acceptor [33,34]. A clear manifestation of the H-bond formation with π-electrons was previously reported by Suzuki et al., wherein they observed 1:1 clusters of benzene with water [35]. Therefore, we can safely assume that the high furfuryl alcohol selectivity may be due to H-a bond formation between water molecules and the π-electrons from the furan ring of 2-furaldehyde, which hampers the adsorption of this aromatic ring on the surface of the catalyst. The carbonyl oxygen atom would be likely to coordinate to the ruthenium surface and would thus receive the hydrogen atoms already present at the surface of the nanoparticle, thus yielding furfuryl alcohol almost exclusively in water.
Catalytic hydrogenation of 2-furaldehyde was also studied in other green solvents such as ethanol, methanol, tetrahydrofuran, and acetonitrile. However, furfuryl alcohol selectivity was inferior although Ru(0)-montmorillonite 3 was prepared in ethanol. Catalyst 3 demonstrated optimal performance at 40 °C in water. At lower temperatures, relatively poor conversion of 2-furaldehyde was observed, whereas the catalyst selectivity suffered at higher temperatures. Montmorillonite 1 was also tested on its own (after activation under exactly the same conditions) and found no conversion of furfural into furfuryl alcohol. Similarly, furfural hydrogenation without a catalyst did not proceed at all, since no substrate conversion was observed. Commercial production of furfuryl alcohol is achieved in gas-phase using Cu-based catalysts at higher temperatures (>100 °C), which results in hydrogenolysis reactions leading to the formation of bi-products [3]. However, hydrogenation of 2-furaldehyde with Ru(0)-montmorillonite 3 occurs at a mild temperature (40 °C), which affords furfuryl alcohol selectively (>99%). In order to determine the turnover number, 25 μL of 2-furaldehyde was added at regular intervals (12 h), until the catalyst lost its activity. The total volume of the substrate used was 125 μL (Table 1).
Ru(0)-hectorite 3 was regenerated by ethanol washing (1 mL, 3×) followed by its activation. The regenerated Ru(0)-hectorite 3 could selectively transform 2-furaldehyde into furfuryl alcohol. However, the reaction takes a longer time with recycled nanoparticles. Ruthenium leaching was determined by combining the washings of six consecutive runs followed by their analysis via ICP-MS. No correction was applied since the interfering iron peak was not observed. Overall, less than 1% leaching was observed with respect to original ruthenium loading. Transmission electron microscopy of the catalyst after five runs shows an overall increase in the size of ruthenium nanoparticles (2.4 ± 0.49 nm, Figure S3). The possible reason for catalyst deactivation can tentatively be attributed to the reduction in surface to volume ratio of nanoparticles. Moreover, slight leaching during each run may also have contributed to the catalyst deactivation.

3. Experimental Section

3.1. Syntheses

Sodium montmorillonite (1) was prepared according to the Reinholdt method [36]. By following the Lagaly method [37], cation exchange capacity (CEC) determined was 1.21 mEq g−1. The dimeric complex [Me(C6H4)Pri)2Ru2Cl4] was prepared according to a procedure reported by Smith and co-workers [38].

3.1.1. Preparation of the Ru(II)-Containing Montmorillonite 2

A clear yellow solution was obtained by dissolving [Me(C6H4)Pri)2Ru2Cl4] (91 mg, 0.15 mmol) in argon-saturated de-ionized water. After rigorous 1 h stirring, a pH of the solution was adjusted to 8 with sodium hydroxide (0.1 N). The filtered solution was then poured into 1 g of degassed sodium montmorillonite 1. For the next 6 h, the suspension was stirred at 20 °C to give Ru(II)-containing montmorillonite 2, which was then filtered off and dried in vacuo.

3.1.2. Preparation of the Ru(0)-Containing Montmorillonite 3

In a stainless-steel autoclave, a suspension of the yellow Ru(II)-containing montmorillonite 2 (5 mg, 0.00148 mmol Ru) in absolute ethanol (1 mL) was stirred magnetically at 100 °C under 1 MPa hydrogen pressure for 15 h. After cooling and pressure release, Ru(0)-containing montmorillonite 3 was obtained as a black material. Standard Schlenk technique was employed to remove ethanol under an inert atmosphere followed by the addition of de-ionized and degassed water (1 mL). The catalyst 3 was then activated for 15 h under 1 MPa hydrogen pressure.

3.2. Catalysis

In a stainless-steel autoclave, 2-furaldehyde (25 μL) was introduced to a suspension (1 mL) of Ru(0)-containing montmorillonite 3. At 40 °C, a constant hydrogen pressure (1 MPa) was maintained with intensive stirring for 12 h. Then, after releasing pressure, the catalyst activity and product selectivity was analyzed. In order to determine the turnover number, 2-furaldehyde (25 μL) was added at regular intervals until the catalyst lost its activity. For the recycled and regenerated Ru(0)-containing montmorillonite 3, this same catalytic procedure was followed.

3.3. Analysis

An Agilent 1260 series Ultra Performance LC system with the column ZORBAX SB-C18 (4.6 × 150 mm, 5 μm), both obtained from Agilent (Santa Clara, CA, USA), was used for the analysis of samples. The mobile phase comprised MeOH (A) and water (B), by using the following general solvent gradient system: 5–95% A from 0 to 20 min, 95% A from 20 min to 25 min, followed by 5% A from 25.1 min to 30 min. The flow rate was set at 1.0 mL min−1, and the column temperature was maintained at 25 °C. Quantitative ruthenium analysis in 3 and the catalyst leaching was determined by using Agilent 7700x type inductively coupled plasma mass spectrometry (ICP-MS).

4. Conclusions

Montmorillonite supported ruthenium nanoparticles were found to efficiently catalyze the hydrogenation of 2-furaldehyde in a benign aqueous medium. The best results were achieved under 1 MPa hydrogen pressure at 40 °C (conversion 100%, selectivity >99%, TON = 1165). These nanoparticles are recyclable and reusable.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/11/1/66/s1, Figure S1: SEM micrographs showing the morphology of Ru(0)-montmorillonite 3 prepared in ethanol, Figure S2: SEM micrographs showing the morphology of Ru(0)-montmorillonite 3 prepared in water, Figure S3: TEM micrographs of Ru(0)-montmorillonite 3 after the deactivation, Table S1: Amount of ruthenium confirmed by ICP-MS.

Author Contributions

Conceptualization, F.-A.K. and H.R.E.-S.; methodology, F.-A.K., Y.W. and M.A.; validation, F.-A.K. and B.S.; quantitative analysis, S.Y. and N.N.; catalyst characterization, M.U. and A.A.I. (XRD and BET), B.S. (TEM, Image J), softwares, M.U., M.A. and B.S.; Visualization, I.H.E.A.; Resources, I.H.E.A.; writing—original draft preparation, F.-A.K., B.S., M.A., A.A.I. and H.R.E.-S.; writing—review and editing, I.H.E.A., F.-A.K., H.R.E.-S. and B.S.; funding acquisition, I.H.E.A. All authors have read and agreed to the published version of the manuscript.

Funding

The funding support by Taif University Researchers Supporting Project number (TURSP-2020/27), Taif University, Taif, Saudi Arabia is highly appreciated. H.R.E.-S. is very grateful to the Swedish Research links Grant VR 2016–05885 and the Department of Molecular Biosciences, Wenner-Grens Institute, Stockholm University, Sweden, for the financial support.

Data Availability Statement

Data is contained within the article or Supplementary Material file.

Acknowledgments

I.H.E.A. thanks Taif University Researchers Supporting Project number (TURSP-2020/27), Taif University, Taif, Saudi Arabia. M.U., A.A.I. and M.A. are thankful to the research facilities at King Fahd University of Petroleum & Minerals, and Saudi Aramco Chair Programme (ORCP2390).

Conflicts of Interest

The authors state that there are no conflict of interest associated with this study and this manuscript’s publication.

References

  1. Wang, Y.; Zhao, D.; Rodríguez-Padrón, D.; Len, C. Recent Advances in Catalytic Hydrogenation of Furfural. Catalysts 2019, 9, 796. [Google Scholar] [CrossRef] [Green Version]
  2. Si, Z.; Zhang, X.; Wang, C.; Ma, L.; Dong, R. An Overview on Catalytic Hydrodeoxygenation of Pyrolysis Oil and Its Model Compounds. Catalysts 2017, 7, 169. [Google Scholar] [CrossRef] [Green Version]
  3. Mariscal, R.; Maireles-Torres, P.; Ojeda, M.; Sádaba, I.; López Granados, M. Furfural: A renewable and versatile platform molecule for the synthesis of chemicals and fuels. Energy Environ. Sci. 2016, 9, 1144–1189. [Google Scholar] [CrossRef]
  4. Gómez Millán, G.; Sixta, H. Towards the Green Synthesis of Furfuryl Alcohol in A One-Pot System from Xylose: A Review. Catalysts 2020, 10, 1101. [Google Scholar] [CrossRef]
  5. Iriondo, A.; Agirre, I.; Viar, N.; Requies, J. Value-Added Bio-Chemicals Commodities from Catalytic Conversion of Biomass Derived Furan-Compounds. Catalysts 2020, 10, 895. [Google Scholar] [CrossRef]
  6. Fulajtárova, K.; Soták, T.; Hronec, M.; Vávra, I.; Dobročka, E.; Omastová, M. Aqueous phase hydrogenation of furfural to furfuryl alcohol over Pd–Cu catalysts. Appl. Catal. A Gen. 2015, 502, 78–85. [Google Scholar] [CrossRef]
  7. Dai, Y.; Chu, X.; Gu, J.; Gao, X.; Xu, M.; Lu, D.; Wan, X.; Qi, W.; Zhang, B.; Yang, Y. Water-enhanced selective hydrogenation of cinnamaldehyde to cinnamyl alcohol on RuSnB/CeO2 catalysts. Appl. Catal. A Gen. 2019, 582. [Google Scholar] [CrossRef]
  8. Yuan, Q.; Zhang, D.; Van Haandel, L.; Ye, F.; Xue, T.; Hensen, E.J.M.; Guan, Y. Selective liquid phase hydrogenation of furfural to furfuryl alcohol by Ru/Zr-MOFs. J. Mol. Catal. A Chem. 2015, 406, 58–64. [Google Scholar] [CrossRef]
  9. Chen, X.; Zhang, L.; Zhang, B.; Guo, X.; Mu, X. Highly selective hydrogenation of furfural to furfuryl alcohol over Pt nanoparticles supported on g-C3N4 nanosheets catalysts in water. Sci. Rep. 2016, 6, 1–13. [Google Scholar] [CrossRef]
  10. Ramirez-Barria, C.; Isaacs, M.; Wilson, K.; Guerrero-Ruiz, A.; Rodríguez-Ramos, I. Optimization of ruthenium based catalysts for the aqueous phase hydrogenation of furfural to furfuryl alcohol. Appl. Catal. A Gen. 2018, 563, 177–184. [Google Scholar] [CrossRef] [Green Version]
  11. Yang, J.; Ma, J.; Yuan, Q.; Zhang, P.; Guan, Y. Selective hydrogenation of furfural on Ru/Al-MIL-53: A comparative study on the effect of aromatic and aliphatic organic linkers. RSC Adv. 2016, 6, 92299–92304. [Google Scholar] [CrossRef]
  12. Bagnato, G.; Figoli, A.; Ursino, C.; Galiano, F.; Sanna, A. A novel Ru–polyethersulfone (PES) catalytic membrane for highly efficient and selective hydrogenation of furfural to furfuryl alcohol. J. Mater. Chem. A 2018, 6, 4955–4965. [Google Scholar] [CrossRef]
  13. Taylor, M.J.; Durndell, L.J.; Isaacs, M.A.; Parlett, C.M.A.; Wilson, K.; Lee, A.F.; Kyriakou, G. Highly selective hydrogenation of furfural over supported Pt nanoparticles under mild conditions. Appl. Catal. B Environ. 2016, 180, 580–585. [Google Scholar] [CrossRef]
  14. Selishcheva, S.A.; Smirnov, A.A.; Fedorov, A.V.; Bulavchenko, O.A.; Saraev, A.A.; Lebedev, M.Y.; Yakovlev, V.A. Highly active CuFeAl-containing catalysts for selective hydrogenation of furfural to furfuryl alcohol. Catalysts 2019, 9, 816. [Google Scholar] [CrossRef] [Green Version]
  15. Jiménez-Gómez, C.P.; Cecilia, J.A.; Moreno-Tost, R.; Maireles-Torres, P. Selective Furfural Hydrogenation to Furfuryl Alcohol Using Cu-Based Catalysts Supported on Clay Minerals. Top. Catal. 2017, 60, 1040–1053. [Google Scholar] [CrossRef]
  16. Putrakumar, B.; Seelam, P.K.; Srinivasarao, G.; Rajan, K.; Rajesh, R.; Rao, K.R.; Liang, T. High Performance and Sustainable Copper-Modified Hydroxyapatite Catalysts for Catalytic Transfer Hydrogenation of Furfural. Catalysts 2020, 10, 1045. [Google Scholar] [CrossRef]
  17. Bardestani, R.; Biriaei, R.; Kaliaguine, S. Hydrogenation of Furfural to Furfuryl Alcohol over Ru Particles Supported on Mildly Oxidized Biochar. Catalysts 2020, 10, 934. [Google Scholar] [CrossRef]
  18. Modelska, M.; Binczarski, M.J.; Kaminski, Z.; Karski, S.; Kolesinska, B.; Mierczynski, P.; Severino, C.J.; Stanishevsky, A.; Witonska, I.A. Bimetallic Pd-Au/SiO2 Catalysts for Reduction of Furfural in Water. Catalysts 2020, 10, 444. [Google Scholar] [CrossRef] [Green Version]
  19. Rao, R.; Dandekar, A.; Baker, R.T.K.; Vannice, M.A. Properties of Copper Chromite Catalysts in Hydrogenation Reactions. J. Catal. 1997, 171, 406–419. [Google Scholar] [CrossRef]
  20. Kumar, B.S.; Dhakshinamoorthy, A.; Pitchumani, K. K10 montmorillonite clays as environmentally benign catalysts for organic reactions. Catal. Sci. Technol. 2014, 4, 2378–2396. [Google Scholar] [CrossRef]
  21. Sun, B.; Khan, F.-A.; Süss-Fink, G.; Therrien, B. Metal Catalysts Intercalated in Smectite Clays. In Encapsulated Catalysts; Elsevier: Amsterdam, The Netherlands, 2017; pp. 387–441. [Google Scholar] [CrossRef]
  22. Ross, C.S.; Hendricks, S.B. Minerals of the Montmorillonite Group Their Origin and Relation to Soils and Clays. Geol. Surv. 1945, 71. [Google Scholar] [CrossRef]
  23. Gaur, R.; Mishra, L.; Siddiqi, M.A.; Atakan, B. Ruthenium complexes as precursors for chemical vapor-deposition (CVD). RSC Adv. 2014, 4, 33785–33805. [Google Scholar] [CrossRef] [Green Version]
  24. Cormary, B.; Dumestre, F.; Liakakos, N.; Soulantica, K.; Chaudret, B. Organometallic precursors of nano-objects, a critical view. Dalton Trans. 2013, 42, 12546. [Google Scholar] [CrossRef]
  25. Sun, B.; Khan, F.-A.; Vallat, A.; Süss-Fink, G. NanoRu@hectorite: A heterogeneous catalyst with switchable selectivity for the hydrogenation of quinoline. Appl. Catal. A Gen. 2013, 467, 310–314. [Google Scholar] [CrossRef]
  26. Khan, F.-A.; Vallat, A.; Süss-Fink, G. Highly selective low-temperature hydrogenation of furfuryl alcohol to tetrahydrofurfuryl alcohol catalysed by hectorite-supported ruthenium nanoparticles. Catal. Commun. 2011, 12, 1428–1431. [Google Scholar] [CrossRef]
  27. Khan, F.-A.; Vallat, A.; Süss-Fink, G. Highly selective CC bond hydrogenation in α,β-unsaturated ketones catalyzed by hectorite-supported ruthenium nanoparticles. J. Mol. Catal. A Chem. 2012, 355, 168–173. [Google Scholar] [CrossRef]
  28. Süss-Fink, G.; Khan, F.-A.; Boudon, J.; Spassov, V. Shape- and Size-Selective Preparation of Hectorite-Supported Ruthenium Nanoparticles for the Catalytic Hydrogenation of Benzene. J. Clust. Sci. 2009, 20, 341–353. [Google Scholar] [CrossRef] [Green Version]
  29. Khan, F.-A.; Süss-Fink, G. Superparamagnetic Core-Shell-Type Fe3O4/Ru Nanoparticles as Catalysts for the Selective Hydrogenation of an Unconstrained α,β-Unsaturated Ketone. Eur. J. Inorg. Chem. 2012, 2012, 727–732. [Google Scholar] [CrossRef]
  30. Sun, B.; Carnevale, D.; Süss-Fink, G. Selective N-cycle hydrogenation of quinolines with sodium borohydride in aqueous media catalyzed by hectorite-supported ruthenium nanoparticles. J. Organomet. Chem. 2016, 821, 197–205. [Google Scholar] [CrossRef] [Green Version]
  31. Sun, B.; Süss-Fink, G. Ruthenium-catalyzed hydrogenation of aromatic amino acids in aqueous solution. J. Organomet. Chem. 2016, 812, 81–86. [Google Scholar] [CrossRef]
  32. Stebler-Roethlisberger, M.; Hummel, W.; Pittet, P.A.; Buergi, H.B.; Ludi, A.; Merbach, A.E. Triaqua(benzene)ruthenium(II) and triaqua(benzene)osmium(II): Synthesis, molecular structure, and water-exchange kinetics. Inorg. Chem. 1988, 27, 1358–1363. [Google Scholar] [CrossRef]
  33. Levitt, M.; Perutz, M.F. Aromatic rings act as hydrogen bond acceptors. J. Mol. Biol. 1988, 201, 751–754. [Google Scholar] [CrossRef]
  34. Burley, S.K.; Petsko, G.A. Amino-aromatic interactions in proteins. FEBS Lett. 1986, 203, 139–143. [Google Scholar] [CrossRef] [Green Version]
  35. Suzuki, S.; Green, P.G.; Bumgarner, R.E.; Dasgupta, S.; Goddard, W.A.; Blake, G.A. Benzene Forms Hydrogen Bonds with Water. Science 1992, 257, 942–945. [Google Scholar] [CrossRef]
  36. Reinholdt, M.X. Synthèse en milieu fluoré et caractérisation de phyllosilicates de type montmorillonite. Etude structurale par spectroscopies d’Absorption des Rayons X et de Résonance Magnétique Nucléaire. Ph.D. Thesis, Université de Haute Alsace-Mulhouse, Mulhouse, France, 2006. [Google Scholar]
  37. Ammann, L.; Bergaya, F.; Lagaly, G. Determination of the cation exchange capacity of clays with copper complexes revisited. Clay Miner. 2005, 40, 441–453. [Google Scholar] [CrossRef]
  38. Bennett, M.A.; Huang, T.-N.; Matheson, T.W.; Smith, A.K.; Ittel, S.; Nickerson, W. (η6-Hexamethylbenzene)ruthenium Complexes. In Inorganic Synthesis; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2007; pp. 74–78. [Google Scholar]
Figure 1. Structural model of sodium montmorillonite Na0.4(Al1.6Mg0.4)Si4O10(OH)2 showing exchangeable cations and water molecules in the interlaminar space [22].
Figure 1. Structural model of sodium montmorillonite Na0.4(Al1.6Mg0.4)Si4O10(OH)2 showing exchangeable cations and water molecules in the interlaminar space [22].
Catalysts 11 00066 g001
Scheme 1. Hydrolysis of dimeric ruthenium complex [Me(C6H4)Pri)2Ru2Cl4] in water gives a mixture of aqua complexes, the dicationic tri aqua complex [Me(C6H4)Pri)Ru(H2O)3]2+ being the major product.
Scheme 1. Hydrolysis of dimeric ruthenium complex [Me(C6H4)Pri)2Ru2Cl4] in water gives a mixture of aqua complexes, the dicationic tri aqua complex [Me(C6H4)Pri)Ru(H2O)3]2+ being the major product.
Catalysts 11 00066 sch001
Scheme 2. Na+ cations exchange in montmorillonite 1 (white) against [Me(C6H4)Pri)Ru(H2O)3]2+ cations to give Ru(II)-modified montmorillonite 2 (yellow) and reduction of [Me(C6H4)Pri)Ru(H2O)3]2+ in 2 by hydrogen gives Ru(0)-containing montmorillonite 3.
Scheme 2. Na+ cations exchange in montmorillonite 1 (white) against [Me(C6H4)Pri)Ru(H2O)3]2+ cations to give Ru(II)-modified montmorillonite 2 (yellow) and reduction of [Me(C6H4)Pri)Ru(H2O)3]2+ in 2 by hydrogen gives Ru(0)-containing montmorillonite 3.
Catalysts 11 00066 sch002
Figure 2. XRD pattern (A) and BET isotherm (B) of sodium montmorillonite 1 and ruthenium containing montmorillonite 2 and 3.
Figure 2. XRD pattern (A) and BET isotherm (B) of sodium montmorillonite 1 and ruthenium containing montmorillonite 2 and 3.
Catalysts 11 00066 g002
Figure 3. TEM micrographs with nanoparticle size distribution histograms (inset): Ru(0)-montmorillonite 3 prepared in water (B); and in ethanol (A).
Figure 3. TEM micrographs with nanoparticle size distribution histograms (inset): Ru(0)-montmorillonite 3 prepared in water (B); and in ethanol (A).
Catalysts 11 00066 g003
Table 1. Hydrogenation of 2-furaldehyde in water at 40 °C using Ru(0)-containing montmorillonite 3.
Table 1. Hydrogenation of 2-furaldehyde in water at 40 °C using Ru(0)-containing montmorillonite 3.
Cat. Run2-Furaldehyde Conversion (%)Time (h)TONFurfuryl Alcohol Selectivity (%)Carbon Balance (%)
110012203>9999
210012203>9999
310012201>9898
410012201>9898
510012203>9999
67612154>9976
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Khan, F.-A.; Yaqoob, S.; Nasim, N.; Wang, Y.; Usman, M.; Isab, A.A.; Altaf, M.; Sun, B.; El Azab, I.H.; El-Seedi, H.R. Ruthenium Nanoparticles Intercalated in Montmorillonite (nano-Ru@MMT) Is Highly Efficient Catalyst for the Selective Hydrogenation of 2-Furaldehyde in Benign Aqueous Medium. Catalysts 2021, 11, 66. https://doi.org/10.3390/catal11010066

AMA Style

Khan F-A, Yaqoob S, Nasim N, Wang Y, Usman M, Isab AA, Altaf M, Sun B, El Azab IH, El-Seedi HR. Ruthenium Nanoparticles Intercalated in Montmorillonite (nano-Ru@MMT) Is Highly Efficient Catalyst for the Selective Hydrogenation of 2-Furaldehyde in Benign Aqueous Medium. Catalysts. 2021; 11(1):66. https://doi.org/10.3390/catal11010066

Chicago/Turabian Style

Khan, Farooq-Ahmad, Sana Yaqoob, Nourina Nasim, Yan Wang, Muhammad Usman, Anvarhusein A. Isab, Muhammad Altaf, Bing Sun, Islam H. El Azab, and Hesham R. El-Seedi. 2021. "Ruthenium Nanoparticles Intercalated in Montmorillonite (nano-Ru@MMT) Is Highly Efficient Catalyst for the Selective Hydrogenation of 2-Furaldehyde in Benign Aqueous Medium" Catalysts 11, no. 1: 66. https://doi.org/10.3390/catal11010066

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