Furfural to Furfuryl Alcohol - ACS Publications - American Chemical

Mar 20, 2018 - Figure 1a shows the optimized structure of the 38T cluster ... Figure 2 shows the energy profile and some .... Even though our database...
9 downloads 3 Views 2MB Size
Article Cite This: Inorg. Chem. XXXX, XXX, XXX−XXX

pubs.acs.org/IC

Furfural to Furfuryl Alcohol: Computational Study of the Hydrogen Transfer on Lewis Acidic BEA Zeolites and Effects of Cation Exchange and Tetravalent Metal Substitution Anittha Prasertsab,† Thana Maihom,*,† Michael Probst,‡ Chularat Wattanakit,§ and Jumras Limtrakul∥ †

Department of Chemistry, Faculty of Liberal Arts and Science, Kasetsart University, Kamphaeng Saen Campus, Nakhon Pathom 73140, Thailand ‡ Institute of Ion Physics and Applied Physics, University of Innsbruck, 6020 Innsbruck, Austria § Department of Chemical and Biomolecular Engineering, School of Energy Science and Engineering and ∥Department of Materials Science and Engineering, Vidyasirimedhi Institute of Science and Technology, Rayong 21210, Thailand S Supporting Information *

ABSTRACT: The hydrogen transfer of furfural to furfuryl alcohol with i-propanol as the hydrogen source over cationexchanged Lewis acidic BEA zeolite has been investigated by means of density functional calculations. The reaction proceeds in three steps. First the O−H bond of i-propanol is broken to form a propoxide intermediate. After that, the furylmethoxy intermediate is formed via hydrogen transfer process, and finally furylmethoxy abstracts the proton to form the furfuryl alcohol product. The second step is ratedetermining by requiring the highest activation energy (23.8 kcal/mol) if the reaction takes place on Li-Sn-BEA zeolite. We find that the catalytic activity of various cation-exchanged SnBEA zeolites is in the order Li-Sn-BEA > Na-Sn-BEA > K-SnBEA. The lower activation energy for Li-Sn-BEA compared to Na-Sn-BEA and K-Sn-BEA can be explained by the larger charge transfer from the carbonyl bond to the catalyst, leading to its activation and to the attraction of the hydrogen being transferred. The larger charge transfer in turn is due to the smaller gap between the energies of furfural HOMO and the zeolite LUMO in LiSn-BEA, compared to both Na-Sn-BEA and K-Sn-BEA. In a similar way, we also compare the catalytic activity of tetravalent metal centers (Sn, Zr, and Hf) substituted into BEA and find in the order Zr ≥ Hf > Sn, based on activation energies. Finally we investigate statistically which property of the reactants is a suitable descriptor for an approximative prediction of the reaction rate in order to be able to quickly screen promising catalytic materials for this reaction.

1. INTRODUCTION Due to the depletion and increasing costs of natural resources, biomass is receiving increased attention in the industry. Its utilization provides commercially viable and environmentally friendly alternative routes for producing renewable fuels and also chemicals.1 Furfural, which can be produced by the acidcatalyzed dehydration of xylose,2−4 belongs to a group of important aromatic intermediates for biorefinery. It is a platform molecule that can be converted into a diversity of industrial chemicals and biofuel components through several reactions.5,6 Furfuryl alcohol is widely used for fine chemicals and in the polymer industry. Examples are the production thermostatic resins, synthetic fibers, farm chemicals, and some fine chemicals or chemical intermediates in the manufacture of lysine, vitamin C, and tetrahydrofurfuryl alcohol.1 A number of catalysts have been used for the catalytic hydrogenation of furfural, and especially noteworthy are catalysts with transition metals like Cu, Pt, Pd, Ni, Ru, and Ir.7−10 The hydrogen in these processes is mostly supplied as high-pressure gas phase © XXXX American Chemical Society

H2, however, which causes them to be rather costly and impractical for industrial production. The catalytic transfer hydrogenation (CTH) has been recently developed as an alternative approach for the reduction of the carbonyl groups of furfural.11,12 This process uses hydrogen donors such as alcohols, especially i-propanol as hydrogen sources instead of molecular H2. The reaction is of the Meerwein−Ponndorf−Verley (MPV) type. A number of catalysts have been reported to be active for CTH reactions, particularly organometallic compounds, transition metals, and metal oxides with acid/base properties.13 Besides the metal catalysts, zeolites with Lewis acidic properties are also active in catalyzing CTH reactions.14−19 Corma et al. reported that Sn-Beta is highly active for both Meerwin−Ponndorf−Verley (MPV) reductions and Baeyer− Villiger oxidation reactions.14−16 They also showed the Received: March 20, 2018

A

DOI: 10.1021/acs.inorgchem.8b00741 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 1. 38T cluster model of the cation-exchanged Lewis acidic BEA zeolite: (a) Li-Sn-BEA and (b) its related systems of M1-M2-BEA (M1 = Li, Na, K and M2 = Sn, Zr, Hf). other materials such as metals-organic frameworks36−43 have confirmed its usability. The 6-31G(d,p) basis set is employed for the Si, C, O, Li, Na, K, and H atoms, while Sn, Zr, and Hf atoms are described by the Stuttgart Effective Core Potential basis (ECP).44 The reacting molecules and most of the zeolite structure are allowed to relax during geometry optimizations, while all terminating hydrogen atoms of the zeolite cluster are fixed to their crystal structure to maintain the unique characteristic of the BEA zeolite. The frequency calculations for obtaining the zero point energy (ZPE) corrections and for the identification of minima and transition states were performed at the same level of theory. For the transition states, it was also checked that only one imaginary frequency corresponding to the expected motion of atoms toward the subsequent intermediate or product is presented. The relative energies in this work are electronic energies including the ZPE correction. The natural bond orbital (NBO) method45 was used to determine orbital overlapping. The partial charges and population analysis were obtained from Charge Model 5 (CM5).46 In addition, the rate constants were also calculated by using classical transition-state theory (TST) consistent with the following equation:

capability of Sn-Beta for polarizing the ketone carbonyl group in the MPV reaction between alcohols and ketones. RománLeshkov et. al reported the utilization of Hf-, Zr-, and Sn-Beta zeolites for catalyzing the coupled transfer hydrogenation and etherification reaction of 5-hydroxymethylfurfural (HMF) with primary and secondary alcohols.17 They also found that SnBEA zeolite in combination with Al is able to catalyze the hydrogenation of furfural to γ-valerolactone (GVL) or furfuryl alcohol. Koehle et. al investigated the utilization of Hf-, Zr-, and Sn-Beta zeolites for catalyzing the transfer hydrogenation of furfural via the MPV reduction using a liquid-phase plug-flow microreactor.19 They reported that Hf-Beta exhibits a much higher rate of reaction compared to Sn- and Zr-Beta. Recently, the modification of the Lewis acidic zeolites by cation exchange has also been attempted. These modified zeolites are reported to significantly promote various reactions including glucose isomerization and epimerization and Baeyer−Villiger oxidation reactions.20 To the best of our knowledge, it was not yet considered how cations exchanged on tetravalent metals BEA zeolites perform for the CTH. In the present work we investigate the CTH mechanisms for the conversion of furfural to furfuryl alcohol over such modified Lewis acid zeolites. We use density functional theory (DFT) calculations with the M06-L functional to obtain energies and geometries of reactants, products, intermediates, and transition states along the reaction pathways. We compare the influence of alkali exchange and of the tetravalent metal centers tin (Sn), zirconium (Zr), and hafnium (Hf) and predict the most efficient combinations.

k=

kBT exp(−ΔG # /RT ) h

where k is the rate constants, kB is Boltzmann’s constant, h is Planck’s constant, T is the absolute temperature, R is the universal gas constant, and ΔG# is the difference of free energy between the initial state and the transition state. The rate constants were derived for the reaction temperature of 298.15 K.

3. RESULTS AND DISCUSSION Figure 1a shows the optimized structure of the 38T cluster model of Li-Sn-BEA. The Sn−OH and O−Li bond distances are 2.10 and 1.81 Å, respectively. Li carries a positively charge of +0.76e. The Lewis acidity of Sn-BEA zeolite has been also analyzed by using NBO charges. Most of the density of the lowest unoccupied molecular orbital (LUMO) is located at the Sn site (cf. Figure S1), indicating the ability of the Sn atom to acquire more electron density. We further discuss the mechanisms of the hydrogen transfer from i-propanol to furfural which becomes furfuryl alcohol. It is analogous to the one of the Meerwein−Ponndorf−Verley (MPV) reaction studied previously47 and proceeds through three steps: The O−H bond dissociation in i-propanol, a hydride transfer from propoxide to furfural, and finally the proton abstraction to form furfuryl alcohol. Figure 2 shows the energy profile and some

2. MODEL AND METHODS The BEA zeolite is represented by a finite 38T cluster model that has been cut from the crystal lattice of BEA zeolite21 as shown in Figure 1. This model includes two perpendicular 12-membered rings at the intersection cavity in the channel system of the zeolite. A silicon atom at a T9 site was replaced with metals atom (Sn, Zr, and Hf) to generate the Lewis acid active site (M-BEA zeolites). This model can be hydrolyzed to generate the open Lewis site and to include a hydroxylated metal site (M−OH) and a silanol (Si−OH) group. The alkali cations (Li, Na, and K) are exchanged on the Brønsted site of the silanol group20 as shown in Figure 1. All calculations were performed with the M06-L density functional22 as implemented in the Gaussian 09 code.23 Previous studies of adsorption and reaction mechanisms on zeolites24−35 and also on B

DOI: 10.1021/acs.inorgchem.8b00741 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 2. Energy profiles and some geometric parameters of reactants, intermediates, and transition states involved in the CTH of furfural to furfuryl alcohol on Li-Sn-BEA zeolite (energies in kcal/mol). Op, Of, and Hp are defined as i-propanol hydroxyl oxygen, furfural carbonyl oxygen, and ipropanol hydrogen of center carbon, respectively.

(TS2), the H at the methylene carbon transfers to C1 of the adsorbed furfural molecule to form the furylmethoxy intermediate and acetone. In this step, the length of the C− H bond increases from 1.10 to 1.51 Å, while the C4···H bond lengths are shortened to 1.21 Å (cf. Figure 2). The CM5 population analysis also points toward the hydride character of the H atom with a partial atomic charge of −0.06e. A normalmode analysis of TS2 reveals one imaginary frequency each at 292.0i cm−1. This mode links the cleavage of the C-Hp bond and the forming the C4-Hp bond. The activation energies are 23.8 kcal/mol, higher than the ones found for Sn-BEA,47 which is related to the more stable i-propanol adsorption states. After TS migration, the furylmethoxy intermediate and acetone (Int2) coadsorbs on the zeolite surface with the complexation energy of −30.0 kcal/mol. The acetone molecule is desorbed after this formation with the desorption energy of 8.5 kcal/mol (Int2_D). Subsequently, the furylmethoxy intermediate is hydrogenated into furfuryl alcohol. The transition state (TS3) involves the concerted bond breaking of the O1−H bond and the formation of the Of−H bond. In this process the O1−H distance is lengthened, and the distance of Of−H is contracted (see Figure 2). The resulting transition state is again confirmed by a single imaginary frequency at 752.4i cm−1 which is associated with the breaking of the O1−H bond and the movement of the H to Of. The activation energy is calculated to be 5.2 kcal/mol. Furfuryl alcohol is finally formed on the zeolite surface with an energy of −31.1 kcal/mol with respect to the isolated system (Prod). On the zeolite its hydroxyl group bonds to the Sn metal active site. Its desorption requires 32.4 kcal/mol. We further investigated the effect of the exchange of sodium and potassium instead of lithium on the catalytic activity of LiSn-BEA. As can be seen from Figure 2, the second step of the hydrogen transfer process is clearly rate determining. The steps forming propoxide and subsequently furfuryl alcohol have a lower barrier (cf. Figure 2). Consider only the rate-determining step, we superimpose its energy profiles in Figure 3. Li-Sn-BEA shows the highest catalytic activity and is followed by Na-Sn-

keys geometrical parameters of the optimized structures along the reaction coordinates. Initially, the i-propanol interacts with both Sn and Li cation of the Li-Sn-BEA zeolite via its hydroxide group oxygen (OH). The Sn···Op and Li···Op distances are 2.55 and 2.22 Å, respectively. Along with the formation of the adsorption complexes, due to electron transfer from Op to Sn, the propanol O−H bond length is slightly increased by 0.01 Å. Sn in the Li-Sn-BEA zeolite becomes less positive charge by +0.09e. The adsorption energy is calculated to be −29.2 kcal/ mol. Following the adsorption of i-propanol, the stable propoxide intermediate is formed via the transition state (TS1) where the hydrogen (H) abstraction from the OH group of propanol takes place. In it, the O−H bond of the propanol OH group is broken, and simultaneously the H hydrogen is transferred to the oxygen (O1) on the zeolite active site and forms a new bond with it. This results in a significant elongation of the Op−H bond distance from 0.98 to 1.31 Å, while the O1···H distance is drastically reduced to be 1.16 Å (see Figure 2). The transition state is verified by checking there is only one imaginary frequency (838.2i cm−1), which corresponds to movement along the expected reaction coordinate described above. The activation barrier with respect to the isolated reactants is 15.4 kcal/mol. This is of the same magnitude as the activation energy required for the hydroxyl O−H bond breaking on metal surface catalysts.48,49 From the transition state, the reaction proceeds to the propoxide intermediate (Int1, see Figure 2). The Op−Sn bond is newly formed with a length of 2.12 Å. Other geometrical parameters can be found in Figure 2. The relative energy of this intermediate complex is −18.1 kcal/mol, making it thermodynamically less stable than the adsorption complex. Subsequently, the furfural molecule is adsorbed on the cation site neighboring the propoxide intermediate with an Of···Li distance of 1.87 Å (Int1_Fur). The carbonyl CO bond is lengthened to 1.23 Å, again caused by charge transfer from the carbonyl group compared to cation active zeolite site. The positive charge on Li is reduced to +0.24e. The adsorption energy of furfural is −32.0 kcal/mol. At the transition state C

DOI: 10.1021/acs.inorgchem.8b00741 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 4. Energy profiles for the rate-determining step of the CTH of furfural to furfuryl alcohol on Li-M-BEA zeolites (M = Sn, Zr, Hf). All the energies are reported in kcal/mol.

Figure 3. Energy profile for the rate-determining step of the CTH of furfural to furfuryl alcohol on M-Sn-BEA zeolites (M = Li, Na, K). Energies are given in kcal/mol.

Lewis acidity and, in other words, is the best in attaching electrons from O. Therefore, the O−C bond distances of the propoxide intermediate are in the order Li-Hf-BEA > Li-ZrBEA > Li-Sn-BEA. The structural changes when the transition state of the hydrogen transfer step is formed are very similar for the three metals. The three activation barriers are 23.8, 19.5, and 20.0 kcal/mol for the Sn-, Zr-, and Hf-BEA zeolites, respectively. Again, the higher Lewis acidity of Hf- and Zr-BEA probably leads to its low barrier. Altogether, it might be concluded that two systems Li-Hf-BEA and Li-Zr-BEA behave similarly and are more active than Li-Sn-BEA. We can also calculate the reaction rate from the series of activation energies (see Table S2) using simple transition state theory (TST). They show the forward rates constants k to be in the order LiZr-BEA > Li-Hf-BEA > Li-Sn-BEA, which corresponds to the activation barriers reported above. However, considering the energy barriers for reversing the reaction back to Int1_Fur, the activation barriers are smaller than the forward ones. Moreover, the higher of the TST reaction rate for revering reaction is also obtained which is in the same trend of the cation exchange SnBEA zeolite. The reverse reaction thus seems to be favored over the forward one. This is due to the lower stability of the carbocation-like furylmethoxy intermediate which is coadsorbed with acetone. As a result, this might affect the catalytic effectiveness of these zeolite systems. Hence, the generated acetone molecule in reaction should be desorbed to prevent the reverse reaction as shown in our calculated mechanism. Moreover, one can see that when the Lewis acidity of BEA active site increases, the reverse rate decreases (Table S2). Increasing of active site Lewis acidity is therefore one of strategies to keep the reaction proceeding forward and to improve the catalyst’s performance. It is always of interest to find relationships between electronic properties of the reactants and the barrier height of the subsequent reaction step, in our case the hydrogen transfer process. Such relations can be used to screen other catalytic materials for this reaction without the time-consuming actual transition state calculation. Electronic descriptors could be, for example, cation charge, HOMO and LUMO energies and their gap, or the complexation energy. Even though our database is quite small, we calculated the linear regressions between them and the activation energy. The least-squares fits of the cation charge, the HOMO energy, and the complexation energy

BEA and K-Sn-BEA. The optimized structures for the reaction on Na-Sn-BEA and K-Sn-BEA are given in Figure S1 of the Supporting Information. It can be seen that the adsorption and transition states are similar to the ones found for Li-Sn-BEA. The activation energies for the Li-Sn-BEA, Na-Sn-BEA, and KSn-BEA systems are 23.8, 28.3, and 31.2 kcal/mol, respectively. This can be related to the amount of charge density transferred from the carbonyl group of furfural and the zeolites. We find that the partial charges of the furfural molecule before and after adsorption are +0.16e (Li-Sn-BEA), +0.11e (Na-Sn-BEA), and +0.09e (K-Sn-BEA), respectively, thus appearing in the same order as the activation barriers reported above. The positive charge is largely concentrated in in the furfural carbonyl C of the three molecules Li-Sn-BEA (1.15e), Na-Sn-BEA (1.14e) and K-Sn-BEA (1.13e). We can also relate the difference in the charge transfer to the energy difference between the LUMO of the zeolites and the HOMO of the furfural molecule which are shown in Figure S3. This gap is in the order Li-Sn-BEA < NaSn-BEA < K-Sn-BEA, where a smaller gap means a higher charge transfer. The findings indicate that the charge transfer is intrinsically related to the C−O carbonyl bond weakening in the Li-Sn-BEA system, and, in turn, the carbon atom can better accept the hydrogen transferred from the propoxide intermediate due to a lower activation barrier. The temperature dependence of the reaction rate constants (k) has also been calculated by using simple transition state theory (TST). The k values are listed in Table S2. The rate of the forward reaction is found to be 3−5 orders higher in the case of Li-Sn-BEA zeolite compared to Na-Sn-BEA and also K-Sn-BEA zeolites, respectively. To compare the catalytic activity of the Lewis metal center active site between Li-Sn-BEA, Li-Zr-BEA, and Li-Hf-BEA, we calculated the reaction rate-determining step. We chose to keep Li due to its higher catalytic activity compared to Na and K, as discussed above. The calculated energy profiles of the reaction on all metals substituted zeolites are shown in Figure 4, and the optimized structures for intermediate adsorption and the transition state are supplied in Figure S2 of the Supporting Information. The complexation energies of propoxide and the furfural intermediate (Int1-Fur) with respect to the isolated molecules are −50.1, −53.1, and −55.0 kcal/mol, respectively. The most stable intermediate is Li-Hf-BEA. It has the highest D

DOI: 10.1021/acs.inorgchem.8b00741 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 5. Left panel: Comparison activation energies obtained by standard and orthogonal linear regression from the cation charges. Right panel: Fitting the activation energy to both the cation charge and the complexation energy.

the complexation energy contain partially correct information as was mentioned above (cf. Figure S4). Using these two descriptors indeed results in a much better approximation of the activation energy (Figure 5, right panel).

against the activation energy are plotted as shown in Figure S4 of Supporting Information. One can see that with exception of the complexation energy it is not really justified to combine the two types of metals (Sn, Zr, Hf and Li, Na, K) in one regression. Further, except for the cation charge as an independent variable, the model predicted activation energies for the Zr and Hf compounds are more different than they should be. Such regression calculations assume that the independent variable (for example, the cation charge) is accurate, and the resulting variable (the activation energy) is subject to a normally distributed error. In our case this is clearly not the case since both variables result from complicated quantum chemical calculations and one could even argue that, for example, cation charge or HOMO energy are less rigorously defined than the activation energy. Therefore, a so-called orthogonal (total) least-square fit which allows errors in both variables and minimizes the perpendicular distance of the data to the regression line is more appropriate. We compared the left panel in Figure S4 with such a treatment.50 Since the relative accuracies of the variables are difficult to guess, we normalized their variances to one prior to the calculation. The result (Figure 5, left panel) shows that in this case the difference between both methods is small. The parameters of all regression lines are given in Table 1. Of course, combining several independent variables improves the quality of the model. Given our set of five data points, using more than two makes little sense. Both the cation charge and

4. CONCLUSIONS We applied density functional theory calculations to study the catalytic hydrogen transfer of furfural to furfuryl alcohol by using i-propanol as the hydrogen source over cation exchanged Lewis acid BEA zeolite. The first reaction step is dissociation of the i-propanol O−H bond to form propoxide intermediate inside the zeolite nanocavity. In the second step, this intermediate transfers hydrogen to the methyl carbon of furfural to become a furylmethoxy intermediate. Finally, the furfuryl alcohol product is formed via proton abstraction from this intermediate. The hydrogen transfer step requires the highest activation energy of 23.8 kcal/mol and is considered to be the rate-determining step of the reaction. The catalytic effect of exchanged cations on the rate-determining step is also considered. Li-Sn-BEA (23.8 kcal/mol) has the lowest activation energy compared to Na-Sn-BEA (28.3 kcal/mol) and K-Sn-BEA (31.2 kcal/mol). The catalytic activity of Li-SnBEA zeolites should therefore prevail over Na-Sn-BEA and KSn-BEA. We also compare the catalytic activity of different tetravalent metal centers (Sn, Zr, and Hf). Their catalytic activity is found to be in the order Zr ≥ Hf > Sn. We provide relationships between activation energy on one side and electronic properties or complexation energies of the intermediate on the other side.

Table 1. One-Dimensional Fitting Models Ea [kcal/mol] obtained by fit to cation charge ρCAT [e] cation charge ρCAT [e] (total least-squares) HOMO energy EHOMO [atomic units] complexation energy Ecplx [kcal/mol]



model Ea Ea Ea Ea

= = = =

ASSOCIATED CONTENT

S Supporting Information *

46.59 ρCAT − 13.65 47.98 ρCAT − 14.80 181.9 EHOMO + 64.12 0.99 Ecplx + 73.95

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.8b00741. E

DOI: 10.1021/acs.inorgchem.8b00741 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry



(11) Tan, J. J.; Cui, J. L.; Cui, X. J.; Deng, T. S.; Li, X. Q.; Zhu, Y. L.; Li, Y. W. Graphene-Modified Ru Nanocatalyst for Low-Temperature Hydrogenation of Carbonyl Groups. ACS Catal. 2015, 5, 7379−7384. (12) Gilkey, M. J.; Xu, B. J. Heterogeneous Catalytic Transfer Hydrogenation as an Effective Pathway in Biomass Upgrading. ACS Catal. 2016, 6, 1420−1436. (13) Kobayashi, H.; Matsuhashi, H.; Komanoya, T.; Hara, K.; Fukuoka, A. Transfer Hydrogenation of Cellulose to Sugar Alcohols over Supported Ruthenium Catalysts. Chem. Commun. 2011, 47, 2366−2368. (14) Corma, A.; Domine, M. E.; Nemeth, L.; Valencia, S. Al-Free SnBeta Zeolite as a Catalyst for the Selective Reduction of Carbonyl Compounds (Meerwein-Ponndorf-Verley Reaction). J. Am. Chem. Soc. 2002, 124, 3194−3195. (15) Boronat, M.; Concepcion, P.; Corma, A.; Renz, M.; Valencia, S. Determination of the Catalytically Active Oxidation Lewis acid Sites in Sn-Beta Zeolites, and their Optimisation by the Combination of Theoretical and Txperimental studies. J. Catal. 2005, 234, 111−118. (16) Corma, A.; Domine, M. E.; Valencia, S. Water-Resistant Solid Lewis Acid Catalysts: Meerwein-Ponndorf-Verley and Oppenauer Reactions Catalyzed by Tin-Beta Zeolite. J. Catal. 2003, 215, 294−304. (17) Lewis, J. D.; Van de Vyver, S.; Crisci, A. J.; Gunther, W. R.; Michaelis, V. K.; Griffin, R. G.; Román-Leshkov, Y. A Continuous Flow Strategy for the Coupled Transfer Hydrogenation and Etherification of 5-(Hydroxymethyl)Furfural Using Lewis Acid Zeolites. ChemSusChem 2014, 7, 2255−2265. (18) Antunes, M. M.; Lima, S.; Neves, P.; Magalhaes, A. L.; Fazio, E.; Fernandes, A.; Neri, F.; Silva, C. M.; Rocha, S. M.; Ribeiro, M. F.; Pillinger, M.; Urakawa, A.; Valente, A. One-Pot Conversion of Furfural to Useful Bio-Products in the Presence of a Sn,Al-Containing Zeolite Beta Catalyst Prepared via Post-Synthesis Routes. J. Catal. 2015, 329, 522−537. (19) Koehle, M.; Lobo, R. F. Lewis acidic zeolite Beta catalyst for the Meerwein-Ponndorf-Verley Reduction of Furfural. Catal. Sci. Technol. 2016, 6, 3018−3026. (20) Otomo, R.; Kosugi, R.; Kamiya, Y.; Tatsumi, T.; Yokoi, T. Modification of Sn-Beta Zeolite: Characterization of Acidic/Basic Properties and Catalytic Performance in Baeyer−Villiger Oxidation. Catal. Sci. Technol. 2016, 6, 2787−2795. (21) Higgins, J. B.; LaPierre, R. B.; Schlenker, J. L.; Rohrman, A. C.; Wood, J. D.; Kerr, G. T.; Rohrbaugh, W. J. The Framework Topology of Zeolite Beta. Zeolites 1988, 8, 446−452. (22) Zhao, Y.; Truhlar, D. G. Density Functionals with Broad Applicability in Chemistry. Acc. Chem. Res. 2008, 41, 157−167. (23) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A., Jr.; Peralta, J. E.; Ogliaro, F.; Bearpark, M. J.; Heyd, J.; Brothers, E. N.; Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A. P.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, N. J.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, Ö .; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian, Inc.: Wallingford, CT, USA, 2009. (24) Boekfa, B.; Choomwattana, S.; Khongpracha, P.; Limtrakul, J. Effects of the Zeolite Framework on the Adsorptions and HydrogenExchange Reactions of Unsaturated Aliphatic, Aromatic, and Heterocyclic Compounds in ZSM-5 Zeolite: A Combination of Perturbation Theory (MP2) and a Newly Developed Density Functional Theory (M06-2X) in ONIOM Scheme. Langmuir 2009, 25 (22), 12990−12999. (25) Maihom, T.; Boekfa, B.; Sirijaraensre, J.; Nanok, T.; Probst, M.; Limtrakul, J. Reaction Mechanisms of the Methylation of Ethene with

LUMO of Li-Sn-BEA (Figure S1), optimized structures for the intermediate adsorption and the transition state of the hydrogen transfer process on Na-Sn-BEA, K-Sn-BEA, and Li-Zr-BEA and Li-Hf-BEA (Figure S2), energy levels of the frontier molecular orbitals (FMOs) of zeolites and furfural (Figure S3), relation between activation energy and cation charge, zeolite HOMO energy and propoxidefurfural intermediates complexation energy (Figure S4), relative energies (Table S1), calculated reaction rates (Table S2), and Cartesian coordinates for all optimized species (PDF)

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Thana Maihom: 0000-0002-8180-1218 Chularat Wattanakit: 0000-0003-3419-9874 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This research was supported, in part, by grants from the Thailand Research Fund to T.M. (Grant MRG6180144), the Kasetsart University Research and Development Institute, and the Graduate Program Scholarship from The Graduate School, Kasetsart University to A.P. The authors also acknowledge National e-Science Infrastructure Consortium for providing computing resources that have contributed to the research results reported within this paper.



REFERENCES

(1) Corma, A.; Iborra, S.; Velty, A. Chemical Routes for the Transformation of Biomass into Chemicals. Chem. Rev. 2007, 107, 2411. (2) Zakrzewska, M. E.; Bogel-Lukasik, E.; Bogel-Lukasik, R. Ionic Liquid-Mediated Formation of 5-Hydroxymethylfurfural-A Promising Biomass-Derived Building Block. Chem. Rev. 2011, 111, 397−417. (3) Lange, J.-P.; van der Heide, E.; van Buijtenen, J.; Price, R. Furfural-a Promising Platform for Lignocellulosic Biofuels. ChemSusChem 2012, 5, 150−166. (4) Van Putten, R.-J.; van der Waal, J. C.; de Jong, E.; Rasrendra, C. B.; Heeres, H. J.; de Vries, J. G. Hydroxymethylfurfural, A Versatile Platform Chemical Made from Renewable Resources. Chem. Rev. 2013, 113, 1499−1597. (5) Zeitsch, K. J. The Chemistry and Technology of Furfural and its Many Byproducts, 1st ed.; Elsevier: Amsterdam, 2000; Vol. 13. (6) Elliott, D. C.; Hart, T. R. Catalytic Hydroprocessing of Chemical Models for Bio-oil. Energy Fuels 2009, 23, 631. (7) Rao, R. S.; Baker, R. T. K.; Vannice, M. A. Furfural Hydrogenation over Carbon-Supported Copper. Catal. Lett. 1999, 60, 51. (8) Nagaraja, B. M.; Siva Kumar, V.; Shasikala, V.; Padmasri, A. H.; Sreedhar, B.; Raju, B. D.; Rama Rao, K. S. A Highly Efficient Cu/MgO Catalyst for Vapour Phase Hydrogenation of Furfural to Furfuryl Alcohol. Catal. Commun. 2003, 4, 287−293. (9) Reddy, B. M.; Reddy, G. K.; Rao, K. N.; Khan, A.; Ganesh, I. Silica Supported Transition Metal-Based Bimetallic Catalysts for Vapour Phase Selective Hydrogenation of Furfuraldehyde. J. Mol. Catal. A: Chem. 2007, 265, 276−282. (10) Sitthisa, S.; Pham, T.; Prasomsri, T.; Sooknoi, T.; Mallinson, R. G.; Resasco, D. E. Conversion of Furfural and 2-Methylpentanal on Pd/SiO2 and Pd-Cu/SiO2 Catalysts. J. Catal. 2011, 280, 17−27. F

DOI: 10.1021/acs.inorgchem.8b00741 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry Methanol and Dimethyl Ether over H-ZSM-5: An ONIOM Study. J. Phys. Chem. C 2009, 113, 6654−6662. (26) Maihom, T.; Pantu, P.; Tachakritikul, C.; Probst, M.; Limtrakul, J. Effect of the Zeolite Nanocavity on the Reaction Mechanism of nHexane Cracking: A Density Functional Theory Study. J. Phys. Chem. C 2010, 114, 7850−7856. (27) Wannakao, S.; Warakulwit, C.; Kongpatpanich, K.; Probst, M.; Limtrakul, J. Methane Activation in Gold Cation-Exchanged Zeolites: A DFT Study. ACS Catal. 2012, 2, 986−992. (28) Maihom, T.; Khongpracha, P.; Sirijaraensre, J.; Limtrakul, J. Mechanistic Studies on the Transformation of Ethanol into Ethene over Fe-ZSM-5 Zeolite. ChemPhysChem 2013, 14, 101−107. (29) Maihom, T.; Wannakao, S.; Boekfa, B.; Limtrakul, J. Density Functional Study of the Activity of Gold-Supported ZSM-5 Zeolites for Nitrous Oxide Decomposition. Chem. Phys. Lett. 2013, 556, 217− 224. (30) Maihom, T.; Probst, M.; Limtrakul, J. Density Functional Theory Study of the Dehydrogenation of Ethanol to Acetaldehyde over the Au-Exchanged ZSM-5 Zeolite: Effect of Surface Oxygen. J. Phys. Chem. C 2014, 118, 18564−18572. (31) Wannakao, S.; Maihom, T.; Probst, M.; Limtrakul, J.; Kongpatpanich, K. Porous Materials as a Platform for Highly Uniform Single-Atom Catalysts: Tuning the Electronic Structure for the LowTemperature Oxidation of Carbon Monoxide. J. Phys. Chem. C 2016, 120 (35), 19686−19697. (32) Maihom, T.; Probst, M.; Limtrakul, J. A DFT Study of Tungsten−Methylidene Formation on a W/ZSM-5 Zeolite: The Metathesis Active Site. ChemPhysChem 2015, 16, 3334−3339. (33) Charoenwiangnuea, P.; Maihom, T.; Kongpracha, P.; Sirijaraensre, J.; Limtrakul, J. Adsorption and Decarbonylation of Furfural over H-ZSM-5 Zeolite: a DFT Study. RSC Adv. 2016, 6, 105888−105894. (34) Injongkol, Y.; Maihom, T.; Boekfa, B.; Limtrakul, J.; Choomwattana, S. A Mechanistic Study of Ethanol Transformation into Ethene and Acetaldehyde on an Oxygenated Au-Exchanged ZSM5 zeolite. RSC Adv. 2017, 7, 38052−38058. (35) Injongkol, Y.; Maihom, T.; Treesukul, P.; Sirijaraensre, J.; Boekfa, B.; Limtrakul, J. Theoretical Study on Reaction Mechanisms of Furfural Hydrogenation to Furfuryl Alcohol on Lewis Acidic BEA Zeolites: Effects of Defect Structure and Tetravalent Metals Substitution. Phys. Chem. Chem. Phys. 2017, 19, 24042−24048. (36) Getman, R. B.; Miller, J. H.; Wang, K.; Snurr, R. Q. Metal Alkoxide Functionalization in Metal-Organic Frameworks for Enhanced Ambient Temperature Hydrogen Storage. J. Phys. Chem. C 2011, 115, 2066−2075. (37) Maihom, T.; Wannakao, S.; Boekfa, B.; Limtrakul, J. Production of Formic Acid via Hydrogenation of CO2 over a Copper-AlkoxideFunctionalized MOF: A Mechanistic Study. J. Phys. Chem. C 2013, 117, 17650−17658. (38) Lee, K.; Isley, W. C.; Dzubak, A. L.; Verma, P.; Stoneburner, S. J.; Lin, L. C.; Howe, J. D.; Bloch, E. D.; Reed, D. A.; Hudson, M. R.; Brown, C. M.; Long, J. R.; Neaton, J. B.; Smit, B.; Cramer, C. J.; Truhlar, D. G.; Gagliardi, L. Design of a Metal−Organic Framework with Enhanced Back Bonding for Separation of N2 and CH4. J. Am. Chem. Soc. 2014, 136 (2), 698−704. (39) Raksakoon, C.; Maihom, T.; Probst, M.; Limtrakul, J. Hydration of Carbon Dioxide in Copper-Alkoxide Functionalized Metal−Organic Frameworks: A DFT Study. J. Phys. Chem. C 2015, 119 (7), 3564− 3571. (40) Maihom, T.; Choomwattana, S.; Wannakao, S.; Probst, M.; Limtrakul, J. Ethylene Epoxidation with Nitrous Oxide over Fe-BTC Metal-Organic Frameworks: A DFT Study. ChemPhysChem 2016, 17, 3416−3422. (41) Ketrat, S.; Maihom, T.; Wannakao, S.; Probst, M.; Nokbin, S.; Limtrakul, J. Coordinatively Unsaturated Metal−Organic Frameworks M3(btc)2 (M = Cr, Fe, Co, Ni, Cu, and Zn) Catalyzing the Oxidation of CO by N2O: Insight from DFT Calculations. Inorg. Chem. 2017, 56 (22), 14005−14012.

(42) Pahls, D. R.; Ortuño, M. A.; Winegar, P. H.; Cramer, C. J.; Gagliardi, L. Computational Screening of Bimetal-Functionalized Zr6O8 MOF Nodes for Methane C−H Bond Activation. Inorg. Chem. 2017, 56, 8739−8743. (43) Maihom, T.; Sawangphruk, M.; Probst, M.; Limtrakul, J. A Computational Study of the Catalytic Aerobic Epoxidation of Propylene over the Coordinatively Unsaturated Metal-Organic Framework Fe3(btc)2: Formation of Propylene Oxide and Competing Reactions. Phys. Chem. Chem. Phys. 2018, 20, 6726−6734. (44) Dolg, M.; Stoll, H.; Preuss, H.; Pitzer, R. M. Relativistic and correlation effects for element 105 (hahnium, Ha): a comparative study of M and MO (M = Nb, Ta, Ha) using energy-adjusted ab initio pseudopotentials. J. Phys. Chem. 1993, 97, 5852−5859. (45) Reed, A. E.; Weinhold, F. Natural bond orbital analysis of nearHartree−Fock water dimer. J. Chem. Phys. 1983, 78, 4066−4073. (46) Marenich, A. V.; Jerome, S. V.; Cramer, C. J.; Truhlar, D. G. Charge Model 5: An Extension of Hirshfeld Population Analysis for the Accurate Description of Molecular Interactions in Gaseous and Condensed Phases. J. Chem. Theory Comput. 2012, 8, 527−541. (47) Boronat, M.; Corma, A.; Renz, M. Mechanism of the MeerweinPonndorf-Verley-Oppenauer (MPVO) Redox Equilibrium on Sn- and Zr-Beta Zeolite Catalysts. J. Phys. Chem. B 2006, 110, 21168−21174. (48) Michel, C.; Auneau, F.; Delbecq, F.; Sautet, P. C−H versus O− H Bond Dissociation for Alcohols on a Rh(111) Surface: A Strong Assistance from Hydrogen Bonded Neighbors. ACS Catal. 2011, 1 (10), 1430−1440. (49) Li, H.; Henkelman, G. Dehydrogenation Selectivity of Ethanol on Close-Packed Transition Metal Surfaces: A Computational Study of Monometallic, Pd/Au, and Rh/Au Catalysts. J. Phys. Chem. C 2017, 121, 27504−27510. (50) Fortran 77 Program of the Extended Classical Total Least Squares Algorithm; www.netlib.org/vanhuffel/dtls-doc (accessed February 6, 2018).

G

DOI: 10.1021/acs.inorgchem.8b00741 Inorg. Chem. XXXX, XXX, XXX−XXX