Synergistic Catalytic Mechanism of Acidic-Silanol ... - ACS Publications

Hua-Qing Yanga*, Chang-Wei Hub*. aCollege of Chemical Engineering, Sichuan University, Chengdu, Sichuan, 610065, P.R. China. bKey Laboratory of Green ...
0 downloads 0 Views 1MB Size
Subscriber access provided by Iowa State University | Library

C: Surfaces, Interfaces, Porous Materials, and Catalysis

Synergistic Catalytic Mechanism of Acidic-Silanol and Basic-Alkylamine Bifunctional Groups over SBA-15 Zeolite toward Aldol Condensation Jin-Feng Zhang, Zhao-Meng Wang, Ya-Jing Lyu, Hong Xie, Ting Qi, Zhen-Bin Si, Li-Juan Liu, Hua-Qing Yang, and Chang-Wei Hu J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.8b11941 • Publication Date (Web): 01 Feb 2019 Downloaded from http://pubs.acs.org on February 5, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 13 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Synergistic Catalytic Mechanism of Acidic-Silanol and Basic-Alkylamine Bifunctional Groups over SBA-15 Zeolite toward Aldol Condensation Jin-Feng Zhanga, Zhao-Meng Wanga, Ya-Jing Lyub, Hong Xiea, Ting Qia, Zhen-Bin Sia, Li-Juan Liua, Hua-Qing Yanga*, Chang-Wei Hub* aCollege

of Chemical Engineering, Sichuan University, Chengdu, Sichuan, 610065, P.R. China Laboratory of Green Chemistry and Technology, Ministry of Education, College of Chemistry, Sichuan University, Chengdu, Sichuan, 610064, P.R. China bKey

ABSTRACT: It is still not clear at molecular level how the acidic-silanol and basic-alkyamine bifunctional groups over SBA-15 zeolite exert catalytic functions in aldol-condensation of 5-hydroxymethylfurfural (HMF) with acetone, which plays pivotal roles in the synthesis of renewable fuels from biomass. Here, the catalytic mechanisms of the aldol-condensation (1) of 5-hydroxymethylfurfural (HMF) with acetone to produce 4-[5-(hydroxymethyl)-2-furanyl]-3-buten-2-one (C9H10O3) over the acidic [–SiOH], basic [–RNH2], and cooperative acidic-basic [–SiOH/–RNH2] active sites on the alkylamine-grafted SBA-15 zeolite (SBA-15-RNH2) have been theoretically investigated, combining quantum mechanical and molecular mechanical (QM/MM) calculations, which involves aldol-condensation (2) and dehydration (3). Based on the plausible reaction pathways, the revised mechanism is proposed. Toward the aldol-condensation (2), the cooperative [–SiOH/–RNH2] display better catalytic activity than the single [–RNH2] or [–SiOH], in which the rate-determining step is associated with the formation of alkylenamine from the alkylamine with enol-acetone. Toward the dehydration (3), [– RNH2] shows better catalytic activity than [–SiOH], in which the rate-determining step is concerned with the [1,3]-H shift for yielding the product C9H10O3. Toward the gross reaction (1), over [–SiOH/–RNH2], the optimal reaction pathway includes the enolization of acetone, aldol-condensation of HMF with enol-acetone, and dehydration. Herein, [–SiOH] is answerable for the activation of acetone to enol-acetone, where it severs as a bridge of H-shift. Additionally, [–RNH2] is responsible for the aldolcondensation of HMF with enol-acetone and the subsequent dehydration. These findings may advance the rational design of cooperative acidic-basic catalyst toward the aldol-condensation reaction at low temperature.

1. Introduction With the dwindling of fossil fuels, the synthesis of renewable fuels and valuable chemicals from biomass has become imperative to meet the demand of environment friendly sustainable processes.1,2 The conversion of biomass-derived oxygenates to fuels and chemicals involves the combination and/or coupling of various types of reactions, including hydrolysis, dehydration, C–C hydrogenolysis, C–O hydrogenolysis, aldol condensation, and so on. In this context, aldol condensation reactions are capable to create new C–C bonds, which are typically employed in the pharmaceutical industry and fine chemical production, and hence heavier and more complex molecules have a bright future in the transition from a fossil resources based towards a more sustainable society.3,4 In particular, as a building block platform chemical bridging biomass chemistry and petrochemistry, 5-hydroxymethylfurfural (HMF) which possesses aldehyde groups can react with

acetone via an aldol-condensation reaction to produce 4-[5(hydroxymethyl)-2-furanyl]-3-buten-2-one (C9H10O3), which can be converted intodiesel-quality fuels by total hydrodeoxygenation.5,6 Currently, aldol condensations are industrially mostly catalyzed by using homogeneous base catalysts such as KOH, NaOH, Ca(OH)2, or Na2CO3.7 On the other hand, these reactions can be catalyzed by using homogeneous acids,8 transition metal catalysts,5,9 or bifunctional molecules that contain both an acid and a base functional groups, such as proline, glycine or derivatives of amino acid.10-12 In spite of satisfactory yields achieved, homogeneous catalysis usually faces some difficulties in separation and recirculation from the reaction mixture, which is incompatible with sustainability.13-15 Therefore, to overcome these disadvantages, heterogenous catalysts, as being greener alternatives, have been focused so far on solid acid catalysts, such as, H3PW12O40, TiO2, ZrO2, WOx, SnO2, and mesoporous

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

silicates in the case of aldol condensations.16-18 Among them, the SBA-15 mesoporous silica has drawn a significant interest in the aldol condensation reactions, because of its large pore sizes, high surface areas, and the easy incorporation of different types of active sites.19 Inspired by enzymatic catalyst, the mesoporous silica is usually modified by introducing two distinct functional groups, most commonly an acid and a base, which are spatially separated such that the functional groups do not interfere with each other. Toward the aldol condensation reactions, the bifunctional acid−base active sites, as a cooperative catalyst, exhibits more efficient catalytic performance than the single active species.20-23 Furthermore, the catalytic cooperativity of the acid−base bifunctional mesoporous silica materials can be affected by both their chemical (e.g., types and strengths of acid and base sites, and the relative ratio of acid to base)20,24,25 and physical properties (e.g., the length and distribution of the alkylamines, and pore sizes).26,27 Toward aldol condensation, the activity of the amine groups on a silica material has been found to be promoted by the weakly acidic silanol groups, which intrinsically exist on the silica material surface.28 In addition, weaker acids such as silanols are more useful partners than silica-supported stronger acids at promoting acid−base cooperativity with primary amines.29,30 Recently, Jones and co-workers reported that the catalytic cooperativity increases with the linker length up to aminopropyl (C3), with longer, more flexible linkers (up to C5) providing no additional benefit or the aldol condensation reactions while short linkers (C1 and C2) limit the beneficial amine−silanol cooperativity.23-26 Afterwards, using the molecular dynamics simulations, they also found that C3 propylamine and C4 butylamine linkers display the highest probability of reaction.22 In spite of these efforts in the catalytic cooperativity of the acid−base bifunctional mesoporous silica materials, little is known about the catalytic mechanism at the molecular level toward aldol condensation reaction. To tailor the rational design of future cooperative catalyst for the aldol condensation reaction, it is necessary to understand the catalytic mechanism over acid−base bifunctional mesoporous silica materials. In this paper, we report here the catalytic mechanism for the aldolcondensation of HMF with acetone on aminopropyl–grafted SBA-15 silica surface. The present study ascertains the optimal reaction pathway over acidic-basic [–SiOH/–RNH2] active sites on SBA-15-RNH2 zeolite, obtains the catalytic orientation of acidic silanol ([–SiOH]) and basic amine ([– RNH2]) functional group, and provides a deep insight into the acid–base catalytic cooperativity on SBA-15-RNH2 zeolite toward aldol condensation, which may advance the rational design of cooperative acid-base catalyst toward the aldol-condensation reaction at low temperature.

2. Computational Methods All of the model structures were performed with Materials Studio 7.0 software package.31 A crystalline silica (001) surface, as being representative of SBA-15 zeolite surface, was prepared from the bulk phase of the α-quartz, based on the ripe method of constructing model in literature.22 The bulk parameters of the SBA-15 zeolite

Page 2 of 13

(001) surface model were first optimized according to the values of the database from Materials Studio. The supercell was obtained by a lattice constant of a = 4.910 Å, and a periodically reproduced slab supercell was then applied to simulate the SBA-15 (001) surface. The simulation cell dimensions were fixed to a = b = 19.64 Å, c = 35.17 Å, and α = β = 90º, and γ = 120º, including six Si−O layers and six tetrahedral Si layers. To keep the original hybridization of silicon, hydroxyls were introduced to saturate the cleaved bonds of surface. Afterward, all of the silicon atoms linked with three hydroxyls were removed. To ensure that the simulation box is neutrally charge, the dangling bonds are once again saturated with hydrogen atoms. The top six floors up and down are symmetrical. The vacuum region was set to 25 Å in order to separate the slabs in the direction perpendicular to the surface. The periodic sianol– terminated silica (001) surface consists of repeated siloxane bonds (Figure 1) as a slab structure with a thickness of 2nm. Then, an ethoxy-propylamine was grafted by incorporating on Si atom of the SBA-15 surface. The stoichiometry of the slab is restricted to 112 silicon, 223 oxygen, 5 carbon, and 75 hydrogen atoms, denoted as Si112O223C5NH75.

Figure 1. SBA-15-RNH2 QM/MM model, the acidic surface silanols [–SiOH] and the basic alkylamine [–RNH2] are shown. Atoms shown in yellow, red, green, gray and blue represent Si, O, H, C and N atoms, respectively. Spherical atoms are QM atoms, others are MM atoms.

All of the model structures were computed by using QMERA module, which can combine quantum mechanics (QM) and molecular mechanics (MM) calculations on periodic systems.32 For the QM region, the generalized gradient approximation (GGA) with PBE functional33 was chosen together with the doubled numerical basis set and polarization basis set (DNP),34 using the DMol3 module.35 The size of the DNP basis set is comparable to Gaussian 631G(d,p), and the DNP basis set is more accurate than a Gaussian basis set of the same size.36,37 The core electrons were treated with all electron.34,35 For the MM region, the GULP38 module was applied with the universal force field (UFF).39 The UFF has been successfully employed to study the zeolites.40,41 For the geometric optimization of intermediates and transition states, both all atoms of the QM region (adsorbates and the surrounding atoms of catalytic active site) and all atoms of the uppermost two layers of the SBA-15 support were relaxed, whereas the remaining atoms of the SBA-15

ACS Paragon Plus Environment

Page 3 of 13 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

support were frozen to mimic bulk constraints. A real-space cutoff of 3.5 Å was used, which is sufficient for the accurate evaluation of the energies.42 The forces imposed on each atom were converged to be less than 0.002 Hartree/Å , while the total energy was converged to be less than 1.0×105 Hartree, and the displacement convergence was less than 5×10-3 Å. A Fermi smearing of 0.005 Hartree for orbital occupancy was used to improve the computational performance. The charge of atom was calculated using the Hirshfeld method.43 The electron density difference was computed using the CASTEP module available in Materials Studio 7.0 package.44 For all the reactants, intermediates and products, the optimized geometric structures were verified to have real frequencies. The transformation pathways and transition state structures were calculated using the complete linear synchronous transit and quadratic synchronous transit (LST/QST) method.45 Each transition state was confirmed to have only one imaginary frequency and its vibration mode had the right direction connecting the reactant and product. The efficiency of the catalyst can be determined by the turnover frequency (TOF) of the catalytic cycle. Based on transition state theory (TST), the TOF can be evaluated by eqs. (i) and (ii) proposed by Kozuch et al.,46-49 in which δE (the energetic span50,51 is defined as the energy difference between the summit and trough of the catalytic cycle TOF 

δE kBT  RT e h

 GTDTS  GTDI δE    GTDTS  GTDI  Gr

(i) if TDTS appears after TDI  if TDTS appears before TDI 

(ii) where kB is the Boltzmann constant, T is the absolute temperature and h is Planck's constant. GTDTS and GTDI are the Gibbs free energies of the TOF-determining transition state (TDTS) and the TOF-determining intermediate (TDI) and ΔGr is the global free energy of the whole cycle. The rate constants k(T) have been evaluated over the 300 ~ 400 K temperature range according to conventional transition state theory TST kʹ(T), including tunneling correction κ(T), as mentioned in our previous study.52

3. Results and Discussions For HMF molecules, formyl group(–CHO) and hydroxymethyl group (–CH2OH) can rotate around the C–C single bond. Thus, HMF may exist eight conformers in acetone solution. The geometric structures and relative Gibbs free energy for eight conformers of HMF in acetone solution are shown in Figure S1 from Supporting Information (SI). As depicted in Figure S1, for the eight conformers of HMF, the relative energies increase as HMF– 6 < HMF–2 < HMF–4 < HMF–8 <HMF–3< HMF–1 < HMF–5 [–RNH2] > [–SiOH]. To explore the origin of the catalytic activity difference of active sites ([–SiOH], [–RNH2], and [–SiOH/–RNH2]), the HOMO and LUMO molecular orbitals of active sites are separately analyzed. The HOMO-LUMO gaps are computed to be 5.820, 5.050, and 4.912 eV for [–RNH2], [–SiOH] and [– SiOH/–RNH2], respectively. It is indicated that the catalytic activity correlates inversely with the corresponding HOMOLUMO gap. In short, the higher catalytic activity stems from the narrower HOMO-LUMO gap. As mentioned earlier, toward the activation of acetone, the corresponding crucial transition sates are O2-TS1 and N1-TS1 over [–SiOH] and [–RNH2], respectively. Toward the formation of alkylenamine over [–RNH2], the corresponding derterminant transition states are ON1-TS1 and N1-TS3, which are from enol-acetone in P-ON1 and from acetone through alkylimine in P-N1, respectively. To visualize the interaction of pivotal transition states, the electron density

differences are analyzed. Figure 5 shows the electron density difference contour map of key transition states. In parts A and B of Figure 5, for the contour plots of the activation of acetone over [–SiOH] (O2-TS1) and over [– RNH2] (N1-TS1), an electron density rich region (red color) between six-membered ring of O1*–H1–C1–C2–O2–H2– (O1*) in O2-TS1 is observably wider than that between four-membered ring N1*-C1-O1-H1-(N1*) in N1-TS1. It is indicated that the electron delocalization in O2-TS1 is stronger than that in N1-TS1, which makes the relative energy of O2-TS1 lower than that of N1-TS1. Thereupon, [– SiOH] shows better catalytic activity than [–RNH2] toward the activation of acetone.

Figure 5. Electron density difference contour plots of (A) O2TS1, (B) N1-TS1, (C) ON1-TS1 and (D) N1-TS3 over SBA-15RNH2 zeolite surface. The red and blue colors represent the increasing and decreasing electron densities, respectively.

In parts C and D of Figure 5, for the contour plots of the formation of alkylenamine over [–RNH2] from enol-acetone in P-ON1 (ON1-TS1) and from acetone through alkylimine in P-N1 through alkylimine (N1-TS3), an electron density rich region (red color) between four-membered ring of N1*–C1–O1–H1–(N1*) in ON1-TS1 is prominently wider than that between four-membered ring N1*-C1-C2-H1(N1*) in N1-TS3. It is inferred that the electron delocalization in ON1-TS1 is dramatically stronger than in N1-TS3, which induces the relative energy of ON1-TS1 lower than that of N1-TS3. Therefore, the formation of alkylenamine from enol-acetone in P-ON1 is superior to that from acetone through alkylimine in P-N1.

4. Conclusions The reaction mechanisms of the aldol-condensation of HMF with acetone catalyzed over the acidic [–SiOH], basic [–RNH2], and cooperative acidic-basic [–SiOH/–RNH2] active sites on SBA-15-RNH2 zeolite have been theoretically investigated. The following conclusions can be drawn from the present investigation. Toward the aldol-condensation (2) of HMF with acetone to C9H12O4, the catalytic activity of active sites decreases as [–SiOH/–RNH2] > [–RNH2] > [–SiOH], which correlates inversely with the corresponding HOMO-LUMO gap. Particularly, the cooperative acidic-basic [–SiOH/–RNH2] exhibit better catalytic activity than the single basic [–RNH2]

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

or single acidic [–SiOH]. Over the acidic-basic [–SiOH/– RNH2], the rate-determining step is concerned with the formation of alkylenamine from the alkylamine with enolacetone. Toward the dehydration reaction (3) of C9H12O4 to C9H10O3, the basic [–RNH2] exhibits better catalytic activity than the acidic [–SiOH]. Over the basic [–RNH2], the ratedetermining step is associated with the [1,3]-H shift for the regenerating the catalyst and yielding the product C9H10O3. Toward the gross reaction (1) of HMF with acetone to C9H10O3, over the acidic-basic [–SiOH/–RNH2], the optimal reaction pathway involves three sequential reaction stages: the enolization of acetone, the aldol-condensation of HMF with enol-acetone, and the dehydration. Herein, the acidic [–SiOH] is responsible for the activation of acetone to enolacetone, where it behaves as a bridge of H-shift. In contrast, the basic [–RNH2] is in charge of the aldol-condensation of HMF with enol-acetone and the subsequent dehydration.



ASSOCIATED CONTENT

Supporting Information The Supporting Information containing is available free of charge via the Internet at The Supporting Information is available free of charge on the ACS Publication website at DOI: 10.1021/xxxx.xxxxxxx. Sum of electronic energies (E, Hartree), and relative energies (Er, kJ mol-1) of various species with respect to the ground reactants. The geometric structures and schematic energy diagrams for the relevant reactions. The entire space models for all species. The evaluation for the sensitivity of force field. The validation of the solvent effect. Arrhenius plots of calculated rate constants for the crucial reaction steps. (PDF) The geometric structures of various species. (MOL Files)



AUTHOR INFORMATION

Corresponding Author * H.-Q. Yang; e-mail: [email protected]; Fax: 86 28 85415608; Telephone: 86 28 85415608 * C.-W. Hu; e-mail: [email protected]; Fax: 86 28 85411105; Telephone: 86 28 85411105

ORCID Jin-Feng Zhang: 0000-0002-9065-626X Zhao-Meng Wang: 0000-0003-0953-6744 Ya-Jing Lyu: 0000-0002-3625-549X Hong Xie: 0000-0002-7638-5862 Ting Qi: 0000-0003-3042-0156 Zhen-Bin Si: 0000-0003-3415-7244 Li-Juan Liu: 0000-0002-7770-8096 Hua-Qing Yang: 0000-0002-2985-0389 Chang-Wei Hu: 0000-0002-4094-6605

Author Contributions The manuscript was written through contributions of all authors. J.-F. Zhang is responsible for main of computation and writing, Z.-M. Wang, Yue Wang, Y.-J. Lyu, and L.-J. Liu for part computation and model design, H. Xie, T. Qi, and Z.-B. Si for analysis, H.-Q. Yang for design, analysis, and writing, and C.-W.

Page 10 of 13

Hu for design and revision. All authors have given approval to the final version of the manuscript.

Notes The authors declare no competing financial interest.



ACKNOWLEDGMENT

The authors are grateful for financial support by the National Natural Science Foundation of China (No: 21573154) and the 111 Project (B17030).



REFERENCES

(1) Corma, A.; Iborra, S.; Velty, A. Chemical Routes for the Transformation of Biomass into Chemicals. Chem. Rev. 2007, 107, 2411–2502. (2) Gallezot, P. Conversion of Biomass to Selected Chemical Products. Chem. Soc. Rev. 2012, 41, 1538–1558. (3) Wu, L.; Moteki, T.; Gokhale, A, A.; Flaherty, D, W.; Toste, F. D. Production of Fuels and Chemicals from Biomass: Condensation Reactions and Beyond. Chem. 2016, 1, 32–58. (4) Li, H.; Riisager, A.; Saravanamurugan, S.; Pandey, A.; Sangwan, R. S.; Yang, S.; Luque, R. Carbon-Increasing Catalytic Strategies for Upgrading Biomass into Energy-Intensive Fuels and Chemicals. ACS Catal. 2017, 8, 148–187. (5) Chatterjee, M.; Matsushima, K.; Ikushima Y.; Sato, M.; Yokoyama, T.; Kawanami, H.; Suzuki, T. Production of Linear Alkane via Hydrogenative Ring Opening of a Furfural-Derived Compound in Supercritical Carbon Dioxide. Green Chem. 2010, 12, 779–782. (6) Chheda, J. N.; Dumesic, J. A. An Overview of Dehydration, Aldol-Condensation and Hydrogenation Processes for Production of Liquid Alkanes from Biomass-Derived Carbohydrates. Catal. Today 2007, 123, 59–70. (7) West, R. M.; Liu, Z. Y.; Peter, M.; Gärtner, C. A.; Dumesic, J. A. Carbon–Carbon Bond Formation for Biomass-Derived Furfurals and Ketones by Aldol Condensation in a Biphasic System. J. Mol. Catal. A: Chem. 2008, 296, 18–27. (8) Cota, I.; Medina, F.; Sueiras, J. E.; Tichit, D. 1,5,7-Triazabicyclo [4.4.0] Dec-5-ene (TBD) an Efficient Homogeneous Catalyst for Aldol Condensation Reactions. Study of the Catalyst Recovery and Reusability Using CO2. Tetrahedron Lett. 2011, 52, 385–387. (9) Mukherjee, S.; Yang, J. W.; Hoffmann, S.; List, B. Asymmetric Enamine Catalysis. Chem. Rev. 2007, 107, 5471–5569. (10) Notz, W.; Tanaka, F.; Barbas, C. F. Enamine-Based Organocatalysis with Proline and Diamines: The Development of Direct Catalytic Asymmetric Aldol, Mannich, Michael, and Diels-Alder Reactions. Acc. Chem. Res. 2004, 37, 580–591. (11) Schmid, M. B.; Zeitler, K.; Gschwind, R. M. The Elusive Enamine Intermediate in Proline-Catalyzed Aldol Reactions: NMR Detection, Formation Pathway, and Stabilization Trends. Angew. Chem., Int. Ed. 2010, 49, 4997–5003. (12) Bahmanyar, S.; Houk, K. N. Transition States of AmineCatalyzed Aldol Reactions Involving Enamine Intermediates: Theoretical Studies of Mechanism, Reactivity, and Stereoselectivity. J. Am. Chem. Soc. 2001, 123, 11273-11283. (13) Tsubogo, T.; Ishiwata, T.; Kobayashi, S. Asymmetric Carbon– Carbon Bond Formation under Continuous-Flow Conditions with Chiral Heterogeneous Catalysts. Angew. Chem., Int. Ed. 2013, 52, 6590–6604. (14) Finelli, F. G.; Miranda, L. S. M.; de Souza, R. O. M. Expanding the Toolbox of Asymmetric Organocatalysis by Continuous-Flow Process. Chem. Commun. 2015, 51, 3708–3722. (15) Vural Gürsel, I.; Noël, T.; Wang, Q.; Hessel, V. Separation/Recycling Methods for Homogeneous Transition Metal Catalysts in Continuous Flow. Green Chem. 2015, 17, 2012–2026. (16) Dedsuksophon, W.; Faungnawakij, K.; Champreda, V.; Laosiripojana, N. Hydrolysis/Dehydration/Aldol-

ACS Paragon Plus Environment

Page 11 of 13 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Condensation/Hydrogenation of Lignocellulosic Biomass and Biomass-Derived Carbohydrates in the Presence of Pd/WO3ZrO2 In a Single Reactor. Bioresour. Technol. 2011, 102, 2040– 2046. (17) Mirjalili, B. F.; Zaghaghi, Z. H3PW12O40/SiO2: An Eco-friendly Alternative for the Stereo-, Regio- and Chemoselective Claisen-Schmidt Condensation. J. Chin. Chem. Soc-TAIP, 2008, 55, 694-697. (18) Matsuhashi, H.; Miyazaki, H.; Kawamura, Y.; Nakamura, H.; Arata, K. Preparation of a Solid Superacid of Sulfated Tin Oxide with Acidity Higher than That of Sulfated Zirconia and Its Applications to Aldol Condensation and Benzoylation. Chem. Mater. 2001, 13, 3038–3042. (19) Zhao, D. Y.; Feng, J. L.; Huo, Q. S.; Melosh, N.; Fredrickson, G. H.; Chmelka, B. F.; Stucky, G. D. Triblock Copolymer Syntheses of Mesoporous Silica with Periodic 50 to 300 Angstrom Pores. Science 1998, 279, 548–552. (20) Lauwaert, J.; De Canck, E.; Esquivel, D.; Thybaut, J. W.; Van Der  Voort, P.; Marin, G. B. Silanol-Assisted Aldol Condensation on Aminated Silica: Understanding the Arrangement of Functional Groups. ChemCatChem 2014, 6, 255–264. (21) Lauwaert, J.; De Canck, E.; Esquivel, D.; Van Der Voort, P.; Thybaut, J. W.; Marin, G. B. Effects of Amine Structure and Base Strength on Acid–Base Cooperative Aldol Condensation. Catal. Today 2015, 246, 35–45. (22) Kim, K. C.; Moschetta, E. G.; Jones, C. W.; Jang, S. S. Molecular Dynamics Simulations of Aldol Condensation Catalyzed by Alkylamine-Functionalized Crystalline Silica Surfaces. J. Am. Chem. Soc. 2016, 138, 7664–7672. (23) Collier, V. E.; Ellebracht, N. C.; Lindy, G. I.; Moschetta, E. G.; Jones, C. W. Kinetic and Mechanistic Examination of Acid– Base Bifunctional Aminosilica Catalysts in Aldol and Nitroaldol Condensations. ACS Catal. 2016, 6, 460–468. (24) Brunelli, N. A.; Venkatasubbaiah, K.; Jones, C. W. Cooperative Catalysis with Acid–Base Bifunctional Mesoporous Silica: Impact of Grafting and Co-condensation Synthesis Methods on Material Structure and Catalytic Properties. Chem. Mater. 2012, 24, 2433–2442. (25) Lauwaert, J.; Moschetta, E. G.; Van Der Voort, P.; Thybaut, J. W.; Jones, C. W.; Marin, G. B. Spatial Arrangement and Acid Strength Effects on Acid–Base Cooperatively Catalyzed Aldol Condensation on Aminosilica Materials. J. Catal. 2015, 325, 19–25. (26) Brunelli, N. A.; Didas, S. A.; Venkatasubbaiah, K.; Jones, C. W. Tuning Cooperativity by Controlling the Linker Length of Silica-Supported Amines in Catalysis and CO2 Capture. J. Am. Chem. Soc. 2012, 134, 13950–13953. (27) Brunelli, N. A.; Jones, C. W. Tuning Acid–Base Cooperativity to Create Next Generation Silica-Supported Organocatalysts. J. Catal. 2013, 308, 60–72. (28) Kubota, Y.; Yamaguchi, H.; Yamada, T.; Inagaki, S.; Sugi, Y.; Tatsumi, T. Further Investigations on the Promoting Effect of Mesoporous Silica on Base-Catalyzed Aldol Reaction. Top Catal. 2010, 53, 492–499. (29) Margelefsky, E. L.; Zeidan, R. K.; Davis, M. E. Cooperative Catalysis by Silica-Supported Organic Functional Groups. Chem. Soc. Rev. 2008, 37, 1118–1126. (30) Sankaranarayanapillai, S.; Sreekumar, S.; Gomes, J.; Grippo, A.; Arab, G. E.; Head-Gordon, M.; Toste, F. D.; Bell, A. T. Catalytic Upgrading of Biomass-Derived Methyl Ketones to Liquid Transportation Fuel Precursors by an Organocatalytic Approach. Angew. Chem., Int. Ed. 2015, 54, 4673–4677. (31) Materials Studio, Version 7.0, Accelrys Software Inc., San Diego, CA, 2013. (32) Allis, D. G.; Prokhorova, D. A.; Korter, T. M. Solid-State Modeling of the Terahertz Spectrum of the High Explosive HMX. J. Phys. Chem. A 2006, 110, 1951–1959.

(33) Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865– 3868. (34) Delley, B. An All-Electron Numerical Method for Solving the Local Density Functional for Polyatomic Molecules. J. Chem. Phys. 1990, 92, 508–517. (35) Delley, B. From Molecules to Solids with the DMol3 Approach. J. Chem. Phys. 2000, 113, 7756–7764. (36) Benedek, N. A.; Snook, I. K.; Latham, K.; Yarovsky, I. Application of Numerical Basis Sets to Hydrogen Bonded Systems: A Density Functional Theory Study. J. Chem. Phys. 2005, 122, 144102. (37) Inada, Y.; Orita, H. Efficiency of Numerical Basis Sets for Predicting the Binding Energies of Hydrogen Bonded Complexes: Evidence of Small Basis Set Superposition Error Compared to Gaussian Basis Sets. J. Comput. Chem. 2008, 29, 225–232. (38) Gale, J. D. GULP: Capabilities and Prospects. Z. Kristallogr. 2005, 220, 552–554. (39) Rappé, A. K.; Casewit, C. J.; Colwell, K. S.; Goddard III, W. A.; Skiff, W. M. UFF, a Full Periodic Table Force Field for Molecular Mechanics and Molecular Dynamics Simulations. J. Am. Chem. Soc. 1992, 114, 10024–10035. (40) Nie, X. W.; Janik, M. J.; Guo, X. W.; Liu, X.; Song, C. S. Reaction Mechanism of Tert-Butylation of Phenol with Tert-Butyl Alcohol over H-β zeolite: An ONIOM Study. Catal. Today 2011, 165, 120–128. (41) Patet, R. E.; Caratzoulas, S.; Vlachos, D. G. Adsorption in Zeolites Using Mechanically Embedded ONIOM Clusters. Phys. Chem. Chem. Phys. 2016, 18, 26094–26106. (42) Lu, Z. S.; Xu, G. L.; He, C. Z.; Wang, T. X.; Yang, L.; Yang, Z. X.; Ma, D. W. Novel Catalytic Activity for Oxygen Reduction Reaction on MnN4 Embedded Graphene: A Dispersion-Corrected Density Functional Theory Study. Carbon. 2015, 84, 500–508. (43) Bultinck, P.; Van Alsenoy, C.; Ayers, P. W.; Carbo-Dorca, R. Critical Analysis and Extension of The Hirshfeld Atoms in Molecules. J. Chem. Phys. 2007, 126, 144111. (44) Seidl, A.; Görling, A.; Vogl, P.; Majewski, J. A.; Levy, M. Generalized Kohn-Sham Schemes and the Band-Gap Problem. Phys. Rev. B 1996, 53, 3764–3774. (45) Govind, N.; Petersen, M.; Fitzgerald, G.; King-Smith, D.; Andzelm, J. A Generalized Synchronous Transit Method for Transition State Location. Comp. Mater. Sci. 2003, 28, 250– 258. (46) Kozuch, S.; Shaik, S. A Combined Kinetic-Quantum Mechanical Model for Assessment of Catalytic Cycles: Application to Cross-Coupling and Heck Reactions. J. Am. Chem. Soc. 2006, 128, 3355–3365. (47) Kozuch, S.; Shaik, S. Kinetic-Quantum Chemical Model for Catalytic Cycles: The Haber-Bosch Process and the Effect of Reagent Concentration. J. Phys. Chem. A 2008, 112, 6032– 6041. (48) Uhe, A.; Kozuch, S.; Shaik, S. Automatic Analysis of Computed Catalytic Cycles. J. Comput. Chem. 2011, 32, 978–985. (49) Kozuch, S.; Skaik, S. How to Conceptualize Catalytic Cycles? The Energetic Span Model. Acc. Chem. Res. 2011, 44, 101–110. (50) Amatore, C.; Jutand, A. Mechanistic and Kinetic Studies of Palladium Catalytic Systems. J. Organomet. Chem. 1999, 576, 254–278. (51) Kozuch, S. Steady State Kinetics of Any Catalytic Network: Graph Theory, the Energy Span Model, the Analogy between Catalysis and Electrical Circuits, and the Meaning of “Mechanism”. ACS Catal. 2015, 5, 5242–5255. (52) Ren, L.-K.; Yang, H.-Q.; Hu, C.-W. Theoretical Study of the Catalytic Oxidation Mechanism of 5-Hydroxymethylfurfural to 2,5-Diformylfuran by PMo-Containing Keggin Heteropolyacid. Catal. Sci. Technol. 2016, 6, 3776–3787. (53) Kandel, K.; Althaus, S. M.; Peeraphatdit, C.; Kobayashi, T.; Trewyn, B. G.; Pruski, M.; Slowing, I. I. Substrate Inhibition in

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

the Heterogeneous Catalyzed Aldol Condensation: A Mechanistic Study of Supported Organocatalysts. J. Catal. 2012, 291, 63–68. (54) Kandel, K.; Althaus, S. M.; Peeraphatdit, C.; Kobayashi, T.; Trewyn, B. G.; Pruski, M.; Slowing, I. I. Solvent-Induced Reversal of Activities between Two Closely Related Heterogeneous Catalysts in the Aldol Reaction. ACS Catal. 2013, 3, 265–271. (55) Sakthivel, K.; Notz, W.; Bui, T.; Barbas Carlos F. Barbas III, C. F. Amino Acid Catalyzed Direct Asymmetric Aldol Reactions: A Bioorganic Approach to Catalytic Asymmetric Carbon-Carbon Bond-Forming Reactions. J. Am. Chem. Soc. 2001, 123, 5260– 5267.

Page 12 of 13

(56) Yutthalekha, T.; Suttipat, D.; Salakhum, S.; Thivasasith, A.; Nokbin, S.; Limtrakul, J.; Wattanakit, C. Aldol Condensation of Biomass-Derived Platform Molecules over Amine-Grafted Hierarchical FAU-Type Zeolite Nanosheets (Zeolean) Featuring Basic Sites. Chem. Commun. 2017, 53, 12185– 12188. (57) Cueto, J.; Faba, L.; Díaz, E,; Ordóñez, S. Performance of Basic Mixed Oxides for Aqueous-Phase 5-HydroxymethylfurfuralAcetone Aldol Condensation. Appl. Catal. B: Environ. 2017, 201, 221-231.

ACS Paragon Plus Environment

Page 13 of 13 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

TOC Graphic

ACS Paragon Plus Environment

13