DFT Modeling of the Adsorption of Trimethylphosphine Oxide at the Internal and External Surfaces of Zeolite MFIClick to copy article linkArticle link copied!
Abstract
The characterization of the acidity of zeolites allows a direct correlation with their catalytic activity. To this end, probe molecules are utilized to obtain a ranking of acid strengths. Trimethylphosphine oxide (TMPO) is a widely used probe molecule, which allows the sensing of solid acids by using 31P NMR. We have performed calculations based on the density functional theory to investigate the Brønsted acid (BA) sites in zeolite MFI by adsorbing TMPO as a probe molecule. We have considered the substitution of silicon at the T2 site by aluminum, both at the internal cavity and at the external surface. The different acid strengths observed in the zeolite MFI when probed by TMPO (very strong, strong, and weak) may depend on the basicity of the centers sharing the acid proton. If the proton lies between the TMPO and one of the framework oxygen atoms binding the Al, the acidity is strong. When the framework oxygen atom is not directly binding the Al, it is less basic and a shortening of the TMPO–H distance is observed, causing an acid response of very strong. Finally, if two TMPO molecules share the proton, the TMPO–H distance elongates, rendering a weak acid character.
1 Introduction
Figure 1
Figure 1. Representation of the BA sites at the internal (bottom, left panel) and at the external (bottom, right panel) surfaces. One of the two pentasil layers that form the slab is identified by a black-line rectangle (top, right panel). The Al-substituted T2-site (light blue ball) with the proton (white ball) at the O1 position (red ball) are shown. The rest of the O atoms and silanol OH groups were deleted for a better view; Si atoms are represented by orange sticks. A molecule of trimethylphosphine oxide is shown in the bottom-right corner; H is represented in white, C in gray, P in brown, and O in red.
2 Computational Methods
2.1 Classification of the Acid Strength
Figure 2
Figure 2. Correlation of the O(P)–H distances and the 31P chemical shifts using Zheng et al. data (black squares linked with black lines). (11) The experimental classification of the acid sites of the zeolite MFI according to the TMPO 31P chemical shift is indicated by vertical dashed blue lines. (10, 12) The experimental classification is used to extrapolate the expected O(P)–H distances from its interception with the theoretical curve (indicated by red circles); the corresponding distance values are written above the horizontal dashed black lines. The spectrum of O(P)–H distances is divided into three zones taking the middle points between the extrapolated O(P)–H distances. These zones are shaded alternately in light gray and white, corresponding to (top) weak acids, (center) strong acids, and (bottom) very strong acids; the limits of each range are indicated by the red numbers at the left-hand side of the graph.
3 Results and Discussion
3.1 Adsorption of One TMPO Molecule
internal BA | external BA | |||
---|---|---|---|---|
Figure 3aa | Figure 3b | Figure 3c | Figure 3d | |
O(P)–H | 1.060 (2.000) | 1.045 (2.000) | 1.066 (2.000) | 1.060 (2.000) |
O(Al)–H | 1.459 (0.975) | 1.483 (0.975) | 1.429 (0.975) | 1.465 (0.975) |
O(P)–O(Al) | 2.518 (2.975) | 2.526 (2.225) | 2.492 (2.975) | 2.522 (2.225) |
P–O(P)b | 1.556 | 1.561 | 1.556 | 1.556 |
The values before the structural optimization are presented within parentheses.
The optimized P−O(P) distance in gas phase is 1.494 Å.
Figure 3
Figure 3. Representation of the interaction of a single TMPO molecule with (a, b) the internal and (c, d) the external BA sites after local optimization; H in white, C in gray, P in brown, O in red, Al in light blue, and Si represented by orange sticks. All the framework O atoms (except the protonated one) and silanol OH groups were deleted for an enhanced view. Related structural values are presented in Table 1.
internal BA | external BA | |||
---|---|---|---|---|
Figure 3aa | Figure 3b | Figure 3c | Figure 3d | |
O(P)–H | 1.06 ± 0.05 | 1.05 ± 0.05 | 1.08 ± 0.08 | 1.07 ± 0.06 |
O(Al)–H | 1.5 ± 0.1 | 1.6 ± 0.2 | 1.5 ± 0.1 | 1.5 ± 0.1 |
O(P)–O(Al) | 2.51 ± 0.09 | 2.6 ± 0.1 | 2.53 ± 0.09 | 2.53 ± 0.09 |
P–O(P) | 1.56 ± 0.03 | 1.57 ± 0.03 | 1.57 ± 0.04 | 1.57 ± 0.03 |
Average and standard deviation over the last 2.5 ps of MD simulation.
Figure 4
3.2 Adsorption of Two TMPO Molecules
Figure 5
Figure 5. Representation of the adsorption of two TMPO molecules on (a) the internal and (b) the external BA sites after local optimization. H in white, C in gray, P in brown, O in red, Al in light blue, and Si represented by orange sticks. All the framework O atoms (except the protonated one) and silanol OH groups were deleted for an enhanced view. Related structural values are presented in Table 3.
internal BA | external BA | |
---|---|---|
Figure 5aa | Figure 5b | |
O(P1)–H | 1.016 (2.000) | 1.033 (2.000) |
O(P2)–H | 2.406 (2.000) | 2.746 (2.000) |
O(P1)–O(P2) | 2.868 (2.462) | 2.908 (2.462) |
O(Al)–H | 1.635 (0.975) | 1.541 (0.975) |
O(P1)–O(Al) | 2.547 (2.372) | 2.544 (2.372) |
P–O(P1) | 1.561 | 1.555 |
P–O(P2) | 1.504 | 1.499 |
The values before the structural optimization are presented within parentheses.
Figure 6
internal BA | external BA | |
---|---|---|
Figure 5aa | Figure 5b | |
O(P1)–H | 1.05 ± 0.04 | 1.07 ± 0.06 |
O(P2)–H | 5.0 ± 0.3 | 8.7 ± 0.2 |
O(P1)–O(P2) | 4.7 ± 0.4 | 8.7 ± 0.2 |
O(Al)–H | 1.5 ± 0.1 | 1.5 ± 0.1 |
O(P1)–O(Al) | 2.52 ± 0.09 | 2.5 ± 0.1 |
P–O(P1) | 1.57 ± 0.03 | 1.56 ± 0.02 |
P–O(P2) | 1.51 ± 0.02 | 1.52 ± 0.01 |
Average and standard deviation over the last 2.5 ps of simulation.
3.3 Adsorption of Three TMPO Molecules
Figure 7
Figure 7. Representation of the adsorption of three TMPO molecules on (a) the internal and (b) the external BA sites after local optimization. H in white, C in gray, P in brown, O in red, Al in light blue, and Si represented by orange sticks. All the framework O atoms (except the protonated one) and silanol OH groups were deleted for an enhanced view. The interaction between one of the nonprotonated TMPO and the Al atom is represented by a stick connecting O(P2) to the Al atom. Related structural values are presented in Table 5.
internal BA | external BA | |
---|---|---|
Figure 7aa | Figure 7b | |
O(P1)–H | 1.131 (2.000) | 1.154 (2.000) |
O(P2)–H | 2.878 (2.000) | 3.134 (2.000) |
O(P3)–H | 3.125 (2.000) | 3.599 (2.000) |
O(P1)–O(P2) | 3.548 (3.570) | 4.034 (3.572) |
O(P1)–O(P3) | 3.473 (3.230) | 4.144 (3.227) |
O(P2)–Al | 1.993 (3.080) | 2.033 (3.073) |
O(Al)–H | 1.250 (0.975) | 1.232 (0.975) |
O(P1)–O(Al) | 2.381 (2.627) | 2.379 (2.627) |
P–O(P1) | 1.533 | 1.531 |
P–O(P2) | 1.511 | 1.511 |
P–O(P3) | 1.501 | 1.496 |
The values before the structural optimization are presented within parentheses.

Figure 8
Figure 8. Charge difference isosurfaces with values of 0.005 bohr–3 calculated from eq 1. (a) Internal BA site; (b) external BA site. The structures correspond to those shown in Figure 7. H in white, C in gray, P in brown, O in red, Al in light blue, and Si represented by orange sticks. All the framework O atoms (except the protonated one) and silanol OH groups were deleted for an enhanced view.
Figure 9
internal BA configurationsa | ||||
---|---|---|---|---|
O(P2)–Al | O(Al)–Al | O(P1)–H | acid classc | |
a | 2.099 | 1.853 | 1.146 | weak |
bb | 1.993 | 1.846 | 1.131 | weak |
c | 2.149 | 1.839 | 1.107 | weak |
d | 1.988 | 1.852 | 1.137 | weak |
e | 2.245 | 1.822 | 1.093 | strong |
f | 2.002 | 1.844 | 1.116 | weak |
external BA configurationsa | ||||
---|---|---|---|---|
O(P2)–Al | O(Al)–Al | O(P1)–H | acid classc | |
a | 3.050 | 1.793 | 1.050 | strong |
bb | 2.033 | 1.842 | 1.154 | weak |
c | 5.554 | 1.812 | 1.053 | strong |
d | 3.637 | 1.785 | 1.059 | strong |
e | 3.059 | 1.789 | 1.047 | strong |
f | 2.059 | 1.829 | 1.100 | weak |
internal BA | external BA | |
---|---|---|
Figure 7aa | Figure 7b | |
O(P1)–H | 1.2 ± 0.1 | 1.05 ± 0.05 |
O(P2)–H | 3.7 ± 0.2 | 6.5 ± 0.3 |
O(P3)–H | 5.6 ± 0.5 | 8.8 ± 0.5 |
O(P1)–O(P2) | 4.9 ± 0.2 | 6.6 ± 0.3 |
O(P1)–O(P3) | 6.3 ± 0.5 | 9.5 ± 0.5 |
O(P2)–Al | 2.0 ± 0.1 | 6.3 ± 0.2 |
O(Al)–H | 1.3 ± 0.2 | 1.5 ± 0.1 |
O(P1)–O(Al) | 2.46 ± 0.09 | 2.52 ± 0.09 |
P–O(P1) | 1.55 ± 0.03 | 1.56 ± 0.02 |
P–O(P2) | 1.52 ± 0.02 | 1.53 ± 0.02 |
P–O(P3) | 1.50 ± 0.02 | 1.51 ± 0.03 |
Average and standard deviation over the last 2.5 ps of simulation.
3.4 Full Deprotonation of the Brønsted Acid Site
Figure 10
Figure 10. Representation of (a, c) TMPOH+ and (b, d) (TMPO)2H+ at more than 5 Å away from the BA site after local optimization. (a, b) Internal surface; (c, d) external surface. H in white, C in gray, P in brown, O in red, and Si represented by orange sticks. All the framework O atoms (except the protonated one) and silanol OH groups were deleted for an enhanced view. Related structural values are presented in Table 8.
internal BA | external BA | |||
---|---|---|---|---|
TMPOH+ (Figure 10a) | (TMPO)2H+ (Figure 10b) | TMPOH+ (Figure 10c) | (TMPO)2H+ (Figure 10d) | |
local optimizationa | ||||
O(P1)–H | 0.979 | 0.996 | 1.005 | 1.072 |
O(P2)–H | − | 2.167 | − | 1.431 |
P–O(P1) | 1.594 | 1.600 | 1.576 | 1.570 |
P–O(P2) | − | 1.515 | − | 1.522 |
O(Si)–H | 2.802 | − | 1.708 | − |
MD simulationb | ||||
O(P1)–H | 1.00 ± 0.04 | 1.11 ± 0.08 | 1.00 ± 0.02 | 1.2 ± 0.1 |
O(P2)–H | − | 1.4 ± 0.1 | − | 1.3 ± 0.1 |
P–O(P1) | 1.59 ± 0.04 | 1.57 ± 0.03 | 1.60 ± 0.05 | 1.55 ± 0.03 |
P–O(P2) | − | 1.53 ± 0.02 | − | 1.54 ± 0.03 |
O(Si)–H | 1.9 ± 0.4 | − | 2.3 ± 0.5 | − |
4 Conclusions
Acknowledgment
The authors acknowledge the use of the UCL Legion High Performance Computing Facility (Legion@UCL), and associated support services, in the completion of this work. Via our membership of the UK’s HEC Materials Chemistry Consortium, which is funded by EPSRC (EP/L000202), this work used the ARCHER UK National Supercomputing Service. This work was also performed using the computational facilities of the Advanced Research Computing@Cardiff (ARCCA) Division, Cardiff University. We acknowledge UCL and the UK Engineering and Physical Sciences Research Council (Grant EP/K009567/2) for funding. Nora H. de Leeuw thanks the Royal Society for an Industry Fellowship. All data created during this research is openly available from the University of Cardiff Research Portal at http://dx.doi.org/10.17035/d.2016.0009738222.
References
This article references 41 other publications.
- 1Sandoval-Díaz, L.-E.; González-Amaya, J.-A.; Trujillo, C.-A. General Aspects of Zeolite Acidity Characterization Microporous Mesoporous Mater. 2015, 215, 229– 243 DOI: 10.1016/j.micromeso.2015.04.038Google Scholar1General aspects of zeolite acidity characterizationSandoval-Diaz, Luis-Ernesto; Gonzalez-Amaya, Jhon-Alex; Trujillo, Carlos-AlexanderMicroporous and Mesoporous Materials (2015), 215 (), 229-243CODEN: MIMMFJ; ISSN:1387-1811. (Elsevier Inc.)A review; this paper describes the most used techniques in the detn. of the acid property in zeolite materials. Two families of techniques, namely spectrometric (IR, NMR) and adsorption-desorption methods (calorimetry, TPD) are considered. Typical exptl. conditions, schematics of equipment setup, and rules for mol. probe selection are shown. Selected expts. that make use of these methods are briefly discussed, and their most relevant results are presented. This review shows some of the possibilities that can be used to "measure" acidity in zeolites. It is also shown that the interpretation of such results could be deceiving in certain situations, esp. because acidity and catalytic performance are not synonyms and acidity is inherently a relative term.
- 2Vermeiren, W.; Gilson, J.-P. Impact of Zeolites on the Petroleum and Petrochemical Industry Top. Catal. 2009, 52, 1131– 1161 DOI: 10.1007/s11244-009-9271-8Google Scholar2Impact of Zeolites on the Petroleum and Petrochemical IndustryVermeiren, W.; Gilson, J.-P.Topics in Catalysis (2009), 52 (9), 1131-1161CODEN: TOCAFI; ISSN:1022-5528. (Springer)A review. The general features of zeolites that led to their widespread use in oil refining and petrochem. are highlighted as well as the details of their impact on selected processes. The anal. of the catalyst market and the position of zeolites therein is a good indication of their strategic importance. Zeolites have brought many disruptive changes to these fields (e.g. FCC). They impacted also these industries in an equally important way, although more subtle, by incremental improvement of processes. The new and vast challenges facing oil refining and petrochem. as well as the managed transition to sustainable environmental benign transport fuel industries and chem. industries will require creative science and technologies. Zeolites offer the basis of many of these technol. solns. provided efficient and balanced cooperations between industry and academia are further developed.
- 3Weitkamp, J. Zeolites and Catalysis Solid State Ionics 2000, 131, 175– 188 DOI: 10.1016/S0167-2738(00)00632-9Google Scholar3Zeolites and catalysisWeitkamp, J.Solid State Ionics (2000), 131 (1,2), 175-188CODEN: SSIOD3; ISSN:0167-2738. (Elsevier Science B.V.)This review with 37 refs. covers the fundamentals of zeolite materials science and their application as catalysts. After a brief introduction into their structures, the most important parameters are discussed which allow the prepn. of an almost infinite variety of zeolitic materials tailored for a given catalytic application. Zeolites are solid acids, and the chem. nature, the d., strength and location of the acid sites in zeolites are discussed. Shape-selective catalysis, which is a unique feature of zeolites, is briefly addressed.
- 4Moliner, M.; Martínez, C.; Corma, A. Multipore Zeolites: Synthesis and Catalytic Applications Angew. Chem., Int. Ed. 2015, 54, 3560– 3579 DOI: 10.1002/anie.201406344Google Scholar4Multipore Zeolites: Synthesis and Catalytic ApplicationsMoliner, Manuel; Martinez, Cristina; Corma, AvelinoAngewandte Chemie, International Edition (2015), 54 (12), 3560-3579CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)A review; in the last few years, important efforts have been made to synthesize so-called "multipore" zeolites, which contain channels of different dimensions within the same cryst. structure. This is a very attractive subject, since the presence of pores of different sizes would favor the preferential diffusion of reactants and products through those different channel systems, allowing unique catalytic activities for specific chem. processes. In this Review we describe the most attractive achievements in the rational synthesis of multipore zeolites, contg. small to extra-large pores, and the improvements reported for relevant chem. processes when these multipore zeolites have been used as catalysts.
- 5Čejka, J.; Wichterlová, B. Acid-Catalyzed Synthesis of Mono- and Dialkyl Benzenes Over Zeolites: Active Sites, Zeolite Topology, and Reaction Mechanisms Catal. Rev.: Sci. Eng. 2002, 44, 375– 421 DOI: 10.1081/CR-120005741Google Scholar5Acid-catalyzed synthesis of mono- and dialkyl benzenes over zeolites: active sites, zeolite topology, and reaction mechanismsCejka, Jiri; Wichterlova, BlankaCatalysis Reviews - Science and Engineering (2002), 44 (3), 375-421CODEN: CRSEC9; ISSN:0161-4940. (Marcel Dekker, Inc.)A review of mechanisms of catalytic reactions leading to mono- and dialkyl benzenes in the presence of zeolite catalysts which are an important part of petrochem. prodn. Emphasis is on the effect of the type of acid sites, zeolite structure, and reaction conditions on the activity and selectivity of these complex reactions, and particularly on the individual products with respect to their iso- vs. n- and ortho-, meta-, and para-isomers. First, descriptions and anal. were made of the structure and properties of acid sites and zeolite pore inner structures as applied in the synthesis of alkyl benzenes. Different structural types of medium- and large-pore zeolite mol. sieves were tested in C6H6 and PhMe alkylation with MeCH:CH2 and Me2CHOH to give cumene and cymenes. Secondly, individual reactions leading to the synthesis of PhPr, xylenes, ethyltoluenes, diethylbenzenes, and transformation of trimethylbenzenes to xylenes from the viewpoint of their reaction mechanism, function of the acid sites, and inner pore geometry of zeolites were described.
- 6Derouane, E. G.; Védrine, J. C.; Pinto, R. R.; Borges, P. M.; Costa, L.; Lemos, M. A. N. D. A.; Lemos, F.; Ribeiro, F. R. The Acidity of Zeolites: Concepts, Measurements and Relation to Catalysis: A Review on Experimental and Theoretical Methods for the Study of Zeolite Acidity Catal. Rev.: Sci. Eng. 2013, 55, 454– 515 DOI: 10.1080/01614940.2013.822266Google Scholar6The acidity of zeolites: Concepts, measurements and relation to catalysis: A review on experimental and theoretical methods for the study of zeolite acidityDerouane, E. G.; Vedrine, J. C.; Pinto, R. Ramos; Borges, P. M.; Costa, L.; Lemos, M. A. N. D. A.; Lemos, F.; Ribeiro, F. RamoaCatalysis Reviews: Science and Engineering (2013), 55 (4), 454-515CODEN: CRSEC9; ISSN:0161-4940. (Taylor & Francis, Inc.)A review. Considered are all aspects of acidity (nature of acid sites, strength, d., etc.) in solid catalysts and in zeolites in particular. After reminding the definition of acidity in liq. and solid acids, the authors emphasized acidity characterization by the most used phys. techniques, such as Hammett's indicator titrn., microcalorimetry of adsorbed probe mols. (ammonia, pyridine or other amines for acidity characterization and CO2 or SO2 for basicity characterization), ammonia or any amine thermodesorption, IR spectroscopy of hydroxyl groups and of several probe mols. adsorbed (ammonia, pyridine, piperidine, amines, CO, H2, etc.), MAS-NMR of 27Al, 29Si, 1H elements and of 1H, 13C, 31P, etc. of adsorbed probe mols., and model catalytic reactions. Modeling the way the acid features of zeolites influence the catalytic activity of these catalysts toward acid-catalyzed reactions (relation between ammonia desorption activation energy values and catalytic activities, reaction mechanism, and kinetics) completes the general anal. of acidity and zeolite chem.
- 7Zheng, A.; Liu, S.-B.; Deng, F. Acidity Characterization of Heterogeneous Catalysts by Solid-State NMR Spectroscopy Using Probe Molecules Solid State Nucl. Magn. Reson. 2013, 55–56, 12– 27 DOI: 10.1016/j.ssnmr.2013.09.001Google Scholar7Acidity characterization of heterogeneous catalysts by solid-state NMR spectroscopy using probe moleculesZheng, Anmin; Liu, Shang-Bin; Deng, FengSolid State Nuclear Magnetic Resonance (2013), 55-56 (), 12-27CODEN: SSNRE4; ISSN:0926-2040. (Elsevier)A review; characterization of the surface acidic properties of solid acid catalysts is a key issue in heterogeneous catalysis. Important acid features of solid acids, such as their type (Bronsted vs. Lewis acid), distribution and accessibility (internal vs. external sites), concn. (amt.), and strength of acid sites are crucial factors dictating their reactivity and selectivity. This short review provides information on different solid-state NMR techniques used for acidity characterization of solid acid catalysts. In particular, different approaches using probe mols. contg. a specific nucleus of interest, such as pyridine-d5, 2-13C-acetone, trimethylphosphine, and trimethylphosphine oxide, are compared. Incorporation of valuable information (such as the adsorption structure, deprotonation energy, and NMR parameters) from d. functional theory (DFT) calcns. can yield explicit correlations between the chem. shift of adsorbed probe mols. and the intrinsic acid strength of solid acids. Methods that combine exptl. NMR data with DFT calcns. can therefore provide both qual. and quant. information on acid sites.
- 8Xu, J.; Zheng, A.; Yang, J.; Su, Y.; Wang, J.; Zeng, D.; Zhang, M.; Ye, C.; Deng, F. Acidity of Mesoporous MoOx/ZrO2 and WOx/ZrO2Materials: A Combined Solid-State NMR and Theoretical Calculation Study J. Phys. Chem. B 2006, 110, 10662– 10671 DOI: 10.1021/jp0614087Google Scholar8Acidity of Mesoporous MoOx/ZrO2 and WOx/ZrO2 Materials: A Combined Solid-State NMR and Theoretical Calculation StudyXu, Jun; Zheng, Anmin; Yang, Jun; Su, Yongchao; Wang, Jiqing; Zeng, Danlin; Zhang, Mingjin; Ye, Chaohui; Deng, FengJournal of Physical Chemistry B (2006), 110 (22), 10662-10671CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)The acidity of mesoporous MoOx/ZrO2 and WOx/ZrO2 materials was studied in detail by multinuclear solid-state NMR techniques as well as DFT quantum chem. calcns. The 1H MAS NMR expts. clearly revealed two different types of strong Bronsted acid sites on both MoOx/ZrO2 and WOx/ZrO2 mesoporous materials, which were able to prontonate adsorbed pyrine-d5 (resulting in 1H NMR signals at chem. shifts in the range 16-19 ppm) as well as adsorbed trimethylphosphine (giving rise to 31P NMR signal at ∼0 ppm). The 13C NMR of adsorbed 2-13C-acetone indicated that the av. Bronsted acid strength of the two mesoporous materials was stronger than that of zeolite HZSM-5 but still weaker than that of 100% H2SO4, which was in good agreement with theor. predictions. The quantum chem. calcns. revealed the detailed structures of the two distinct types of Bronsted acid sites formed on the mesoporous MoOx/ZrO2 and WOx/ZrO2. The existence of both monomer and oligomer Mo (or W) species contg. a Mo-OH-Zr (or W-OH-Zr) bridging OH group was confirmed with the former having an acid strength close to zeolite HZSM-5, with the latter having an acid strength similar to sulfated zirconia. From the authors' NMR exptl. and theor. calcn. results, a possible mechanism is proposed for the formation of acid sites on these mesoporous materials.
- 9Li, S.; Huang, S.-J.; Shen, W.; Zhang, H.; Fang, H.; Zheng, A.; Liu, S.-B.; Deng, F. Probing the Spatial Proximities among Acid Sites in Dealuminated H-Y Zeolite by Solid-State NMR Spectroscopy J. Phys. Chem. C 2008, 112, 14486– 14494 DOI: 10.1021/jp803494nGoogle Scholar9Probing the Spatial Proximities among Acid Sites in Dealuminated H-Y Zeolite by Solid-State NMR SpectroscopyLi, Shenhui; Huang, Shing-Jong; Shen, Wanling; Zhang, Hailu; Fang, Hanjun; Zheng, Anmin; Liu, Shang-Bin; Deng, FengJournal of Physical Chemistry C (2008), 112 (37), 14486-14494CODEN: JPCCCK; ISSN:1932-7447. (American Chemical Society)A comprehensive study has been made to probe the spatial proximities among different acid sites in dealuminated H-Y zeolites modified with various degrees of calcination, steam, and acid treatments by using a variety of different solid-state NMR techniques, including multinuclear MAS NMR and two-dimensional 1H double-quantum (DQ) MAS NMR spectroscopy. The effects of dealumination treatments on the nature, concn., and location of extraframework Al species in H-Y zeolites were followed by 1H DQ MAS NMR of hydroxyl protons in conjunction with 1H, 27Al, and 29Si MAS NMR results. It was found that the extraframework AlOH species (Lewis acid sites) are always in close proximity to the bridging AlOHSi hydroxyls (Bronsted acid sites) on the framework of dealuminated H-Y zeolites prepd. by thermal and hydrothermal treatments, indicating the presence of a Bronsted/Lewis acid synergy effect. However, such an effect is absent in acid-treated H-Y zeolites, as also confirmed by 13C CP/MAS NMR of adsorbed 2-13C-acetone.
- 10Zhao, Q.; Chen, W.-H.; Huang, S.-J.; Wu, Y.-C.; Lee, H.-K.; Liu, S.-B. Discernment and Quantification of Internal and External Acid Sites on Zeolites J. Phys. Chem. B 2002, 106, 4462– 4469 DOI: 10.1021/jp015574kGoogle Scholar10Discernment and Quantification of Internal and External Acid Sites on ZeolitesZhao, Qi; Chen, Wen-Hua; Huang, Shing-Jong; Wu, Yu-Chih; Lee, Huang-Kuei; Liu, Shang-BinJournal of Physical Chemistry B (2002), 106 (17), 4462-4469CODEN: JPCBFK; ISSN:1089-5647. (American Chemical Society)A new methodol. is reported for concurrent qual. and quant. characterization of internal and external acid sites in zeolitic catalysts. H-ZSM-5 zeolites with varied Si/Al ratios have been examd. by solid-state 31P MAS NMR using different adsorbed probe mols., namely trimethylphosphine oxides (TMPO) and tributylphosphine oxide (TBPO), in conjunction with elemental anal. Up to seven distinct 31P resonance peaks at 86, 75, 67, 63, 53, 43, and 30 ppm were identified from the 31P NMR spectra of adsorbed TMPO. The resonance peak at 30 ppm has never been obsd. previously and may be ascribed to mobile TMPO. The peak at 43 ppm is assigned to physisorbed TMPO. The rest of the peaks result from TMPOH+ complexes at Bronsted sites with peaks at higher chem. shifts reflecting acid sites of higher strengths. 31P NMR expts., performed with TMPO and TBPO adsorbed on a mesoporous MCM-41 sample, resp., provide further correlation of the internal and external acid sites. It is concluded that the peaks at 75 and 53 ppm arise exclusively from the internal Bronsted sites, whereas the peaks at 86, 67, and 63 ppm are assocd. with both internal and external acid sites. While the concn. and distribution of internal sites were found to increase with acid strengths, changing the Si/Al ratio of HZSM-5 has nearly no effect on the strength of the external acid sites.
- 11Zheng, A.; Zhang, H.; Lu, X.; Liu, S.-B.; Deng, F. Theoretical Predictions of 31P NMR Chemical Shift Threshold of Trimethylphosphine Oxide Absorbed on Solid Acid Catalysts J. Phys. Chem. B 2008, 112, 4496– 4505 DOI: 10.1021/jp709739vGoogle Scholar11Theoretical Predictions of 31P NMR Chemical Shift Threshold of Trimethylphosphine Oxide Absorbed on Solid Acid CatalystsZheng, Anmin; Zhang, Hailu; Lu, Xin; Liu, Shang-Bin; Deng, FengJournal of Physical Chemistry B (2008), 112 (15), 4496-4505CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)The 31P NMR chem. shifts of adsorbed trimethylphosphine oxide (TMPO) and the configurations of the corresponding TMPOH+ complexes on Bronsted acid sites with varying acid strengths in modeled zeolites have been predicted theor. by means of d. functional theory (DFT) quantum chem. calcns. The configuration of each TMPOH+ complex was optimized at the PW91/DNP level based on an 8T cluster model, whereas the 31P chem. shifts were calcd. with the gauge including AO (GIAO) approach at both the HF/TZVP and MP2/TZVP levels. A linear correlation between the 31P chem. shift of adsorbed TMPO and the proton affinity of the solid acids was obsd., and a threshold for superacidity (86 ppm) was detd. This threshold for superacidity was also confirmed by comparative investigations on other superacid systems, such as carborane acid and heteropolyoxometalate H3PW12O40. In conjunction with the strong correlation between the MP2 and the HF 31P isotropic shifts, the 8T cluster model was extended to more sophisticated models (up to 72T) that are not readily tractable at the GIAO-MP2 level, and a 31P chem. shift of 86 ppm was detd. for TMPO adsorbed on zeolite H-ZSM-5, which is in good agreement with the NMR exptl. data.
- 12Seo, Y.; Cho, K.; Jung, Y.; Ryoo, R. Characterization of the Surface Acidity of MFI Zeolite Nanosheets by 31 P NMR of Adsorbed Phosphine Oxides and Catalytic Cracking of Decalin ACS Catal. 2013, 3, 713– 720 DOI: 10.1021/cs300824eGoogle Scholar12Characterization of the Surface Acidity of MFI Zeolite Nanosheets by 31P NMR of Adsorbed Phosphine Oxides and Catalytic Cracking of DecalinSeo, Yongbeom; Cho, Kanghee; Jung, Younjae; Ryoo, RyongACS Catalysis (2013), 3 (4), 713-720CODEN: ACCACS; ISSN:2155-5435. (American Chemical Society)MFI zeolite nanosheets tailored to 2.5-nm thickness were synthesized using a surfactant-type zeolite structure-directing agent, [C22H45-N+(CH3)2-C6H12-N+(CH3)2-C6H13](Br-)2. The zeolite nanosheets possessed Bronsted acid sites on their external surfaces as well as in the internal micropore walls. The acid strength and concn. was characterized by the 31P NMR signals of the adsorbed trimethylphosphine oxide and tributylphosphine oxide. The 31P NMR investigation identified three types of Bronsted acid sites with different strengths on external surfaces; there were four types inside the micropores. A linear correlation has been established between the no. of the external strongest acid sites and the catalytic activity in decalin cracking for the MFI zeolite catalysts investigated in this work.
- 13Ni, Y.; Sun, A.; Wu, X.; Hai, G.; Hu, J.; Li, T.; Li, G. The Preparation of Nano-Sized H[Zn,Al]ZSM-5 Zeolite and Its Application in the Aromatization of Methanol Microporous Mesoporous Mater. 2011, 143, 435– 442 DOI: 10.1016/j.micromeso.2011.03.029Google ScholarThere is no corresponding record for this reference.
- 14Costa, C.; Dzikh, I. P.; Lopes, J. M.; Lemos, F.; Ribeiro, F. R. Activity–acidity Relationship in Zeolite ZSM-5. Application of Brönsted-Type Equations J. Mol. Catal. A: Chem. 2000, 154, 193– 201 DOI: 10.1016/S1381-1169(99)00374-XGoogle Scholar14Activity-acidity relationship in zeolite ZSM-5. Application of Bronsted-type equationsCosta, C.; Dzikh, I. P.; Lopes, J. M.; Lemos, F.; Ribeiro, F. R.Journal of Molecular Catalysis A: Chemical (2000), 154 (1-2), 193-201CODEN: JMCCF2; ISSN:1381-1169. (Elsevier Science B.V.)In this paper the relation between activity and acidity in a variety of ZSM-5 zeolite catalysts, with different Si/Al ratios and different protonic content, is analyzed and a quant. correlation is obtained. The acid site strength distribution was estd. using temp.-programmed desorption (TPD) of ammonia by applying a digital deconvolution method to the curves. These data were then correlated with exptl. catalytic activity data for the same catalysts towards n-heptane cracking reaction, by means of a Bronsted-type equation similar to the ones used for homogeneous acid catalysis and already used for other zeolites. It can be noticed that the same types of equation that are used for homogeneous acid catalysis also hold for heterogeneous acid catalysis and that the activation energy for ammonia desorption can be used as acid-strength scale for the purpose of correlation with catalytic activity.
- 15Kwok, T. J.; Jayasuriya, K.; Damavarapu, R.; Brodman, B. W. Application of H-ZSM-5 Zeolite for Regioselective Mononitration of Toluene J. Org. Chem. 1994, 59, 4939– 4942 DOI: 10.1021/jo00096a042Google ScholarThere is no corresponding record for this reference.
- 16Wang, Y.; Jian, G.; Peng, Z.; Hu, J.; Wang, X.; Duan, W.; Liu, B. Preparation and Application of Ba/ZSM-5 Zeolite for Reaction of Methyl Vinyl Ether and Methanol Catal. Commun. 2015, 66, 34– 37 DOI: 10.1016/j.catcom.2015.03.012Google Scholar16Preparation and application of Ba/ZSM-5 zeolite for reaction of methyl vinyl ether and methanolWang, Yang; Jian, Guixing; Peng, Zhihong; Hu, Jiaji; Wang, Xin; Duan, Wubiao; Liu, BoCatalysis Communications (2015), 66 (), 34-37CODEN: CCAOAC; ISSN:1566-7367. (Elsevier B.V.)The ZSM-5 zeolite is widely used to catalyze the reactions of methanol to olefins. Herein, we have prepd. the H-ZSM-5 doped with barium (Ba/ZSM-5) using incipient wetness impregnation method. The Ba modified catalysts were used to catalyze a new reaction of methanol with Me vinyl ether to improve the selectivity of ethylene and propylene (C=2 + C=3). The reaction catalyzed by Ba doped H-ZSM-5 shows higher propylene selectivity over H-ZSM-5. The reaction mechanism is discussed.
- 17Zheng, A.; Han, B.; Li, B.; Liu, S.-B.; Deng, F. Enhancement of Brønsted Acidity in Zeolitic Catalysts due to an Intermolecular Solvent Effect in Confined Micropores Chem. Commun. 2012, 48, 6936– 6938 DOI: 10.1039/c2cc32498aGoogle Scholar17Enhancement of Bronsted acidity in zeolitic catalysts due to an intermolecular solvent effect in confined microporesZheng, Anmin; Han, Bing; Li, Bojie; Liu, Shang-Bin; Deng, FengChemical Communications (Cambridge, United Kingdom) (2012), 48 (55), 6936-6938CODEN: CHCOFS; ISSN:1359-7345. (Royal Society of Chemistry)Solid-state NMR and DFT calcn. studies certified the presence of an intermol. solvent effect for mols. confined in microporous zeolite, leading to a notable increase in Bronsted acidity of the solid acid catalyst.
- 18Kresse, G.; Hafner, J. Ab Initio Molecular Dynamics for Liquid Metals Phys. Rev. B: Condens. Matter Mater. Phys. 1993, 47, 558– 561 DOI: 10.1103/PhysRevB.47.558Google Scholar18Ab initio molecular dynamics of liquid metalsKresse, G.; Hafner, J.Physical Review B: Condensed Matter and Materials Physics (1993), 47 (1), 558-61CODEN: PRBMDO; ISSN:0163-1829.The authors present ab initio quantum-mech. mol.-dynamics calcns. based on the calcn. of the electronic ground state and of the Hellmann-Feynman forces in the local-d. approxn. at each mol.-dynamics step. This is possible using conjugate-gradient techniques for energy minimization, and predicting the wave functions for new ionic positions using sub-space alignment. This approach avoids the instabilities inherent in quantum-mech. mol.-dynamics calcns. for metals based on the use of a factitious Newtonian dynamics for the electronic degrees of freedom. This method gives perfect control of the adiabaticity and allows one to perform simulations over several picoseconds.
- 19Kresse, G.; Hafner, J. Ab Initio Molecular-Dynamics Simulation of the Liquid-Metal–amorphous-Semiconductor Transition in Germanium Phys. Rev. B: Condens. Matter Mater. Phys. 1994, 49, 14251– 14269 DOI: 10.1103/PhysRevB.49.14251Google Scholar19Ab initio molecular-dynamics simulation of the liquid-metal-amorphous-semiconductor transition in germaniumKresse, G.; Hafner, J.Physical Review B: Condensed Matter and Materials Physics (1994), 49 (20), 14251-69CODEN: PRBMDO; ISSN:0163-1829.The authors present ab initio quantum-mech. mol.-dynamics simulations of the liq.-metal-amorphous-semiconductor transition in Ge. The simulations are based on (a) finite-temp. d.-functional theory of the 1-electron states, (b) exact energy minimization and hence calcn. of the exact Hellmann-Feynman forces after each mol.-dynamics step using preconditioned conjugate-gradient techniques, (c) accurate nonlocal pseudopotentials, and (d) Nose' dynamics for generating a canonical ensemble. This method gives perfect control of the adiabaticity of the electron-ion ensemble and allows the authors to perform simulations over >30 ps. The computer-generated ensemble describes the structural, dynamic, and electronic properties of liq. and amorphous Ge in very good agreement with expt.. The simulation allows the authors to study in detail the changes in the structure-property relation through the metal-semiconductor transition. The authors report a detailed anal. of the local structural properties and their changes induced by an annealing process. The geometrical, bounding, and spectral properties of defects in the disordered tetrahedral network are studied and compared with expt.
- 20Kresse, G.; Furthmüller, J. Efficiency of Ab-Initio Total Energy Calculations for Metals and Semiconductors Using a Plane-Wave Basis Set Comput. Mater. Sci. 1996, 6, 15– 50 DOI: 10.1016/0927-0256(96)00008-0Google Scholar20Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis setKresse, G.; Furthmuller, J.Computational Materials Science (1996), 6 (1), 15-50CODEN: CMMSEM; ISSN:0927-0256. (Elsevier)The authors present a detailed description and comparison of algorithms for performing ab-initio quantum-mech. calcns. using pseudopotentials and a plane-wave basis set. The authors will discuss: (a) partial occupancies within the framework of the linear tetrahedron method and the finite temp. d.-functional theory, (b) iterative methods for the diagonalization of the Kohn-Sham Hamiltonian and a discussion of an efficient iterative method based on the ideas of Pulay's residual minimization, which is close to an order N2atoms scaling even for relatively large systems, (c) efficient Broyden-like and Pulay-like mixing methods for the charge d. including a new special preconditioning optimized for a plane-wave basis set, (d) conjugate gradient methods for minimizing the electronic free energy with respect to all degrees of freedom simultaneously. The authors have implemented these algorithms within a powerful package called VAMP (Vienna ab-initio mol.-dynamics package). The program and the techniques have been used successfully for a large no. of different systems (liq. and amorphous semiconductors, liq. simple and transition metals, metallic and semi-conducting surfaces, phonons in simple metals, transition metals and semiconductors) and turned out to be very reliable.
- 21Kresse, G.; Furthmüller, J. Efficient Iterative Schemes for Ab Initio Total-Energy Calculations Using a Plane-Wave Basis Set Phys. Rev. B: Condens. Matter Mater. Phys. 1996, 54, 11169– 11186 DOI: 10.1103/PhysRevB.54.11169Google Scholar21Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis setKresse, G.; Furthmueller, J.Physical Review B: Condensed Matter (1996), 54 (16), 11169-11186CODEN: PRBMDO; ISSN:0163-1829. (American Physical Society)The authors present an efficient scheme for calcg. the Kohn-Sham ground state of metallic systems using pseudopotentials and a plane-wave basis set. In the first part the application of Pulay's DIIS method (direct inversion in the iterative subspace) to the iterative diagonalization of large matrixes will be discussed. This approach is stable, reliable, and minimizes the no. of order Natoms3 operations. In the second part, we will discuss an efficient mixing scheme also based on Pulay's scheme. A special "metric" and a special "preconditioning" optimized for a plane-wave basis set will be introduced. Scaling of the method will be discussed in detail for non-self-consistent and self-consistent calcns. It will be shown that the no. of iterations required to obtain a specific precision is almost independent of the system size. Altogether an order Natoms2 scaling is found for systems contg. up to 1000 electrons. If we take into account that the no. of k points can be decreased linearly with the system size, the overall scaling can approach Natoms. They have implemented these algorithms within a powerful package called VASP (Vienna ab initio simulation package). The program and the techniques have been used successfully for a large no. of different systems (liq. and amorphous semiconductors, liq. simple and transition metals, metallic and semiconducting surfaces, phonons in simple metals, transition metals, and semiconductors) and turned out to be very reliable.
- 22Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple Phys. Rev. Lett. 1996, 77, 3865– 3868 DOI: 10.1103/PhysRevLett.77.3865Google Scholar22Generalized gradient approximation made simplePerdew, John P.; Burke, Kieron; Ernzerhof, MatthiasPhysical Review Letters (1996), 77 (18), 3865-3868CODEN: PRLTAO; ISSN:0031-9007. (American Physical Society)Generalized gradient approxns. (GGA's) for the exchange-correlation energy improve upon the local spin d. (LSD) description of atoms, mols., and solids. We present a simple derivation of a simple GGA, in which all parameters (other than those in LSD) are fundamental consts. Only general features of the detailed construction underlying the Perdew-Wang 1991 (PW91) GGA are invoked. Improvements over PW91 include an accurate description of the linear response of the uniform electron gas, correct behavior under uniform scaling, and a smoother potential.
- 23Grimme, S. Semiempirical GGA-Type Density Functional Constructed with a Long-Range Dispersion Correction J. Comput. Chem. 2006, 27, 1787– 1799 DOI: 10.1002/jcc.20495Google Scholar23Semiempirical GGA-type density functional constructed with a long-range dispersion correctionGrimme, StefanJournal of Computational Chemistry (2006), 27 (15), 1787-1799CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)A new d. functional (DF) of the generalized gradient approxn. (GGA) type for general chem. applications termed B97-D is proposed. It is based on Becke's power-series ansatz from 1997 and is explicitly parameterized by including damped atom-pairwise dispersion corrections of the form C6·R-6. A general computational scheme for the parameters used in this correction has been established and parameters for elements up to xenon and a scaling factor for the dispersion part for several common d. functionals (BLYP, PBE, TPSS, B3LYP) are reported. The new functional is tested in comparison with other GGAs and the B3LYP hybrid functional on std. thermochem. benchmark sets, for 40 noncovalently bound complexes, including large stacked arom. mols. and group II element clusters, and for the computation of mol. geometries. Further cross-validation tests were performed for organometallic reactions and other difficult problems for std. functionals. In summary, it is found that B97-D belongs to one of the most accurate general purpose GGAs, reaching, for example for the G97/2 set of heat of formations, a mean abs. deviation of only 3.8 kcal mol-1. The performance for noncovalently bound systems including many pure van der Waals complexes is exceptionally good, reaching on the av. CCSD(T) accuracy. The basic strategy in the development to restrict the d. functional description to shorter electron correlation lengths scales and to describe situations with medium to large interat. distances by damped C6·R-6 terms seems to be very successful, as demonstrated for some notoriously difficult reactions. As an example, for the isomerization of larger branched to linear alkanes, B97-D is the only DF available that yields the right sign for the energy difference. From a practical point of view, the new functional seems to be quite robust and it is thus suggested as an efficient and accurate quantum chem. method for large systems where dispersion forces are of general importance.
- 24Hernandez-Tamargo, C. E.; Roldan, A.; de Leeuw, N. H. A Density Functional Theory Study of the Structure of Pure-Silica and Aluminium-Substituted MFI Nanosheets J. Solid State Chem. 2016, 237, 192– 203 DOI: 10.1016/j.jssc.2016.02.006Google Scholar24A density functional theory study of the structure of pure-silica and aluminium-substituted MFI nanosheetsHernandez-Tamargo, Carlos E.; Roldan, Alberto; de Leeuw, Nora H.Journal of Solid State Chemistry (2016), 237 (), 192-203CODEN: JSSCBI; ISSN:0022-4596. (Elsevier B.V.)The authors present a theor. study of the structural features of silica and aluminum-substituted MFI nanosheets. The authors have analyzed the effects of aluminum substitution on the vibrational properties of silanols as well as the features of protons as counter-ions. The formation of the two-dimensional system did not lead to appreciable distortions within the framework. Moreover, the effects on the structure due to the aluminum dopants were the same in both the bulk and the slab. The principal differences were related to the silanol groups that form hydrogen-bonds with neighboring aluminum-substituted silanols, whereas intra-framework hydrogen-bonds increase the stability of aluminum-substituted silanols toward dehydration. Thus, the authors have complemented previous exptl. and theor. studies, showing the lamellar MFI zeolite to be a very stable material of high crystallinity regardless of its very thin structure.
- 25Blöchl, P. E. Projector Augmented-Wave Method Phys. Rev. B: Condens. Matter Mater. Phys. 1994, 50, 17953– 17979 DOI: 10.1103/PhysRevB.50.17953Google Scholar25Projector augmented-wave methodBlochlPhysical review. B, Condensed matter (1994), 50 (24), 17953-17979 ISSN:0163-1829.There is no expanded citation for this reference.
- 26Kresse, G.; Joubert, D. From Ultrasoft Pseudopotentials to the Projector Augmented-Wave Method Phys. Rev. B: Condens. Matter Mater. Phys. 1999, 59, 1758– 1775 DOI: 10.1103/PhysRevB.59.1758Google Scholar26From ultrasoft pseudopotentials to the projector augmented-wave methodKresse, G.; Joubert, D.Physical Review B: Condensed Matter and Materials Physics (1999), 59 (3), 1758-1775CODEN: PRBMDO; ISSN:0163-1829. (American Physical Society)The formal relationship between ultrasoft (US) Vanderbilt-type pseudopotentials and Blochl's projector augmented wave (PAW) method is derived. The total energy functional for US pseudopotentials can be obtained by linearization of two terms in a slightly modified PAW total energy functional. The Hamilton operator, the forces, and the stress tensor are derived for this modified PAW functional. A simple way to implement the PAW method in existing plane-wave codes supporting US pseudopotentials is pointed out. In addn., crit. tests are presented to compare the accuracy and efficiency of the PAW and the US pseudopotential method with relaxed-core all-electron methods. These tests include small mols. (H2, H2O, Li2, N2, F2, BF3, SiF4) and several bulk systems (diamond, Si, V, Li, Ca, CaF2, Fe, Co, Ni). Particular attention is paid to the bulk properties and magnetic energies of Fe, Co, and Ni.
- 27Ho, K.-M.; Fu, C. L.; Harmon, B. N.; Weber, W.; Hamann, D. R. Vibrational Frequencies and Structural Properties of Transition Metals via Total-Energy Calculations Phys. Rev. Lett. 1982, 49, 673– 676 DOI: 10.1103/PhysRevLett.49.673Google Scholar27Vibrational frequencies and structural properties of transition metals via total-energy calculationsHo, K. M.; Fu, C. L.; Harmon, B. N.; Weber, W.; Hamann, D. R.Physical Review Letters (1982), 49 (9), 673-6CODEN: PRLTAO; ISSN:0031-9007.The vibrational frequencies of selected normal modes can be obtained entirely from 1st principles with use of frozen phonon calcns. which involve the precise evaluation of crystal total energy as a function of lattice displacement. The calcns. allow a detailed anal. of the microscopic mechanism causing phonon anomalies and soft-mode phase transitions. Calcns. for Zr, Nb, and Mo were made using both tight-binding and pseudopotential methods.
- 28Fu, C.-L.; Ho, K.-M. First-Principles Calculation of the Equilibrium Ground-State Properties of Transition Metals: Applications to Nb and Mo Phys. Rev. B: Condens. Matter Mater. Phys. 1983, 28, 5480– 5486 DOI: 10.1103/PhysRevB.28.5480Google Scholar28First-principles calculation of the equilibrium ground-state properties of transition metals: Application to niobium and molybdenumFu, C. L.; Ho, K. M.Physical Review B: Condensed Matter and Materials Physics (1983), 28 (10), 5480-6CODEN: PRBMDO; ISSN:0163-1829.A self-consistent pseudopotential method was used to calc. the equil. ground-state properties of Mo and Nb. The calcd. equil. lattice consts., cohesive energies, and bulk moduli agree with exptl. data.
- 29Quartieri, S.; Arletti, R.; Vezzalini, G.; Di Renzo, F.; Dmitriev, V. Elastic Behavior of MFI-Type Zeolites: 3 – Compressibility of Silicalite and Mutinaite J. Solid State Chem. 2012, 191, 201– 212 DOI: 10.1016/j.jssc.2012.03.039Google Scholar29Elastic behavior of MFI-type zeolites: 3. Compressibility of silicalite and mutinaiteQuartieri, Simona; Arletti, Rossella; Vezzalini, Giovanna; Di Renzo, Francesco; Dmitriev, VladimirJournal of Solid State Chemistry (2012), 191 (), 201-212CODEN: JSSCBI; ISSN:0022-4596. (Elsevier B.V.)We report the results of an in-situ synchrotron X-ray powder diffraction study, performed using silicone oil as "non-penetrating" pressure transmitting medium, of the elastic behavior of three zeolites with MFI-type framework: the natural zeolite mutinaite and two silicalites (labeled A and B) synthesized under different conditions. In mutinaite, no symmetry change is obsd. as a function of pressure; however, a phase transition from monoclinic (P21/n) to orthorhombic (Pnma) symmetry occurs at about 1.0 GPa in the silicalite samples. This phase transition is irreversible upon decompression. The second order bulk moduli of silicalite A and silicalite B, calcd. after the fulfillment of the phase transition, are: K 0=18.2(2) and K 0=14.3 (2) GPa, resp. These values makes silicalite the most compressible zeolite among those up to now studied in silicone oil. The structural deformations induced by HP in silicalite A were investigated by means of complete Rietveld structural refinements, before and after the phase transition, at P amb and 0.9 GPa, resp. The elastic behaviors of the three MFI-type zeolites here investigated were compared with those of Na-ZSM-5 and H-ZSM-5, studied in similar exptl. conditions: the two silicalites - which are the phases with the highest Si/Al ratios and hence the lowest extraframework contents - show the highest compressibility. On the contrary, the most rigid material is mutinaite, which has a very complex extraframework compn. characterized by a high no. of cations and water mols.
- 30Kokotailo, G. T.; Lawton, S. L.; Olson, D. H.; Meier, W. M. Structure of Synthetic Zeolite ZSM-5 Nature 1978, 272, 437– 438 DOI: 10.1038/272437a0Google Scholar30Structure of synthetic zeolite ZSM-5Kokotailo, G. T.; Lawton, S. L.; Olson, D. H.; Meier, W. M.Nature (London, United Kingdom) (1978), 272 (5652), 437-8CODEN: NATUAS; ISSN:0028-0836.The structure of synthetic zeolite ZSM-5 was detd. The crystals are orthorhombic, space group Pnma, with a 20.1, b 19.9, and c 13.4 Å. The unit contents of the Na form are NanAlnSi96-nO192.∼16H2O with n <27 and typically ∼3. The ZSM-5 framework contains 2 intersecting channel systems, 1 sinusoidal parallel to [001] and 1 straight parallel to [010]. The elliptical 10-membered ring openings controlling the channels have an effective diam. between those of zeolite Linde A and faujasite.
- 31Nosé, S. A Unified Formulation of the Constant Temperature Molecular Dynamics Methods J. Chem. Phys. 1984, 81, 511 DOI: 10.1063/1.447334Google Scholar31A unified formulation of the constant-temperature molecular-dynamics methodsNose, ShuichiJournal of Chemical Physics (1984), 81 (1), 511-19CODEN: JCPSA6; ISSN:0021-9606.Three recently proposed const. temp. mol. dynamics methods [N., (1984) (1); W. G. Hoover et al., (1982) (2); D. J. Evans and G. P. Morris, (1983) (2); and J. M. Haile and S. Gupta, 1983) (3)] are examd. anal. via calcg. the equil. distribution functions and comparing them with that of the canonical ensemble. Except for effects due to momentum and angular momentum conservation, method (1) yields the rigorous canonical distribution in both momentum and coordinate space. Method (2) can be made rigorous in coordinate space, and can be derived from method (1) by imposing a specific constraint. Method (3) is not rigorous and gives a deviation of order N-1/2 from the canonical distribution (N the no. of particles). The results for the const. temp.-const. pressure ensemble are similar to the canonical ensemble case.
- 32Nosé, S. Constant Temperature Molecular Dynamics Methods Prog. Theor. Phys. Suppl. 1991, 103, 1– 46 DOI: 10.1143/PTPS.103.1Google Scholar32Constant-temperature molecular-dynamics methodsNose, ShuichiProgress of Theoretical Physics Supplement (1991), 103 (), 1-46CODEN: PTPSEP; ISSN:0375-9687.A review with 62 refs. on how the canonical distribution is realized in simulations based on deterministic dynamical equations. Basic formulations and extensions of two const.-temp. mol. dynamics methods, the constrained and the extended system methods, are discussed.
- 33Bylander, D. M.; Kleinman, L. Energy Fluctuations Induced by the Nosé Thermostat Phys. Rev. B: Condens. Matter Mater. Phys. 1992, 46, 13756– 13761 DOI: 10.1103/PhysRevB.46.13756Google Scholar33Energy fluctuations induced by the Nose thermostatBylander; KleinmanPhysical review. B, Condensed matter (1992), 46 (21), 13756-13761 ISSN:0163-1829.There is no expanded citation for this reference.
- 34Momma, K.; Izumi, F. VESTA 3 for Three-Dimensional Visualization of Crystal, Volumetric and Morphology Data J. Appl. Crystallogr. 2011, 44, 1272– 1276 DOI: 10.1107/S0021889811038970Google Scholar34VESTA 3 for three-dimensional visualization of crystal, volumetric and morphology dataMomma, Koichi; Izumi, FujioJournal of Applied Crystallography (2011), 44 (6), 1272-1276CODEN: JACGAR; ISSN:0021-8898. (International Union of Crystallography)VESTA is a 3D visualization system for crystallog. studies and electronic state calcns. It was upgraded to the latest version, VESTA 3, implementing new features including drawing the external morphpol. of crysals; superimposing multiple structural models, volumetric data and crystal faces; calcn. of electron and nuclear densities from structure parameters; calcn. of Patterson functions from the structure parameters or volumetric data; integration of electron and nuclear densities by Voronoi tessellation; visualization of isosurfaces with multiple levels, detn. of the best plane for selected atoms; an extended bond-search algorithm to enable more sophisticated searches in complex mols. and cage-like structures; undo and redo is graphical user interface operations; and significant performance improvements in rendering isosurfaces and calcg. slices.
- 35Tuckerman, M.; Laasonen, K.; Sprik, M.; Parrinello, M. Ab Initio Molecular Dynamics Simulation of the Solvation and Transport of Hydronium and Hydroxyl Ions in Water J. Chem. Phys. 1995, 103, 150 DOI: 10.1063/1.469654Google Scholar35Ab initio molecular dynamics simulation of the solvation and transport of hydronium and hydroxyl ions in waterTuckerman, M.; Laasonen, K.; Sprik, M.; Parrinello, M.Journal of Chemical Physics (1995), 103 (1), 150-61CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)Charge defects in water created by excess or missing protons appear in the form of solvated hydronium H3O+ and hydroxyl OH- ions. Using the method of ab initio mol. dynamics, we have investigated the structure and proton transfer dynamics of the solvation complexes, which embed the ions in the network of hydrogen bonds in the liq. In our ab initio mol. dynamics approach, the interat. forces are calcd. each time step from the instantaneous electronic structure using d. functional methods. All hydrogen atoms, including the excess proton, are treated as classical particles with the mass of a deuterium atom. For the H3O+ ion we find a dynamic solvation complex, which continuously fluctuates between a (H5O2)+ and a (H9O4)+ structure as a result of proton transfer. The OH- has a predominantly planar fourfold coordination forming a (H9O5)- complex. Occasionally this complex is transformed in a more open tetrahedral (H7H4)- structure. Proton transfer is obsd. only for the more waterlike (H7O4)- complex. Transport of the charge defects is a concerted dynamical process coupling proton transfer along hydrogen bonds and reorganization of the local environment. The simulation results strongly support the structural diffusion mechanism for charge transport. In this model, the entire structure-and not the constituent particles-of the charged complex migrates through the hydrogen bond network. For H3O+, we propose that transport of the excess proton is driven by coordination fluctuations in the first solvation shell (i.e., second solvation shell dynamics). The rate-limiting step for OH- diffusion is the formation of the (H7O4)- structure, which is the solvation state showing proton transfer activity.
- 36Henkelman, G.; Arnaldsson, A.; Jónsson, H. A Fast and Robust Algorithm for Bader Decomposition of Charge Density Comput. Mater. Sci. 2006, 36, 354– 360 DOI: 10.1016/j.commatsci.2005.04.010Google ScholarThere is no corresponding record for this reference.
- 37Sanville, E.; Kenny, S. D.; Smith, R.; Henkelman, G. Improved Grid-Based Algorithm for Bader Charge Allocation J. Comput. Chem. 2007, 28, 899– 908 DOI: 10.1002/jcc.20575Google Scholar37Improved grid-based algorithm for Bader charge allocationSanville, Edward; Kenny, Steven D.; Smith, Roger; Henkelman, GraemeJournal of Computational Chemistry (2007), 28 (5), 899-908CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)An improvement to the grid-based algorithm of Henkelman et al. for the calcn. of Bader vols. is suggested, which more accurately calcs. at. properties as predicted by the theory of Atoms in Mols. The CPU time required by the improved algorithm to perform the Bader anal. scales linearly with the no. of interat. surfaces in the system. The new algorithm corrects systematic deviations from the true Bader surface, calcd. by the original method and also does not require explicit representation of the interat. surfaces, resulting in a more robust method of partitioning charge d. among atoms in the system. Applications of the method to some small systems are given and it is further demonstrated how the method can be used to define an energy per atom in ab initio calcns.
- 38Tang, W.; Sanville, E.; Henkelman, G. A Grid-Based Bader Analysis Algorithm without Lattice Bias J. Phys.: Condens. Matter 2009, 21, 084204 DOI: 10.1088/0953-8984/21/8/084204Google Scholar38A grid-based Bader analysis algorithm without lattice biasTang, W.; Sanville, E.; Henkelman, G.Journal of Physics: Condensed Matter (2009), 21 (8), 084204/1-084204/7CODEN: JCOMEL; ISSN:0953-8984. (Institute of Physics Publishing)A computational method for partitioning a charge d. grid into Bader vols. is presented which is efficient, robust, and scales linearly with the no. of grid points. The partitioning algorithm follows the steepest ascent paths along the charge d. gradient from grid point to grid point until a charge d. max. is reached. In this paper, we describe how accurate off-lattice ascent paths can be represented with respect to the grid points. This improvement maintains the efficient linear scaling of an earlier version of the algorithm, and eliminates a tendency for the Bader surfaces to be aligned along the grid directions. As the algorithm assigns grid points to charge d. maxima, subsequent paths are terminated when they reach previously assigned grid points. It is this grid-based approach which gives the algorithm its efficiency, and allows for the anal. of the large grids generated from plane-wave-based d. functional theory calcns.
- 39Wernet, P.; Nordlund, D.; Bergmann, U.; Cavalleri, M.; Odelius, M.; Ogasawara, H.; Näslund, L. A.; Hirsch, T. K.; Ojamäe, L.; Glatzel, P. The Structure of the First Coordination Shell in Liquid Water Science 2004, 304, 995– 999 DOI: 10.1126/science.1096205Google Scholar39The Structure of the First Coordination Shell in Liquid WaterWernet, Ph.; Nordlund, D.; Bergmann, U.; Cavalleri, M.; Odelius, M.; Ogasawara, H.; Naeslund, L. A.; Hirsch, T. K.; Ojamaee, L.; Glatzel, P.; Pettersson, L. G. M.; Nilsson, A.Science (Washington, DC, United States) (2004), 304 (5673), 995-999CODEN: SCIEAS; ISSN:0036-8075. (American Association for the Advancement of Science)X-ray absorption spectroscopy and x-ray Raman scattering were used to probe the mol. arrangement in the 1st coordination shell of liq. H2O. The local structure is characterized by comparison with bulk and surface of ordinary hexagonal ice Ih and with calcd. spectra. Most mols. in liq. H2O are in 2 H-bonded configurations with 1 strong donor and 1 strong acceptor H bond in contrast to the 4 H-bonded tetrahedral structure in ice. Upon heating from 25° to 90°, 5 to 10% of the mols. change from tetrahedral environments to 2 H-bonded configurations. Findings are consistent with neutron and x-ray diffraction data, and combining the results sets a strong limit for possible local structure distributions in liq. H2O. Serious discrepancies with structures based on current mol. dynamics simulations are obsd.
- 40Ambrosetti, A.; Alfè, D.; Robert, A.; DiStasio, J.; Tkatchenko, A. Hard Numbers for Large Molecules: Toward Exact Energetics for Supramolecular Systems J. Phys. Chem. Lett. 2014, 5, 849– 855 DOI: 10.1021/jz402663kGoogle Scholar40Hard Numbers for Large Molecules: Toward Exact Energetics for Supramolecular SystemsAmbrosetti, Alberto; Alfe, Dario; DiStasio, Robert A.; Tkatchenko, AlexandreJournal of Physical Chemistry Letters (2014), 5 (5), 849-855CODEN: JPCLCD; ISSN:1948-7185. (American Chemical Society)Noncovalent interactions are ubiquitous in mol. and condensed-phase environments, and hence a reliable theor. description of these fundamental interactions could pave the way toward a more complete understanding of the microscopic underpinnings for a diverse set of systems in chem. and biol. In this work, we demonstrate that recent algorithmic advances coupled to the availability of large-scale computational resources make the stochastic quantum Monte Carlo approach to solving the Schrodinger equation an optimal contender for attaining "chem. accuracy" (1 kcal/mol) in the binding energies of supramol. complexes of chem. relevance. To illustrate this point, we considered a select set of seven host-guest complexes, representing the spectrum of noncovalent interactions, including dispersion or van der Waals forces, π-π stacking, hydrogen bonding, hydrophobic interactions, and electrostatic (ion-dipole) attraction. A detailed anal. of the interaction energies reveals that a complete theor. description necessitates treatment of terms well beyond the std. London and Axilrod-Teller contributions to the van der Waals dispersion energy.
- 41Huang, S.-J.; Yang, C.-Y.; Zheng, A.; Feng, N.; Yu, N.; Wu, P.-H.; Chang, Y.-C.; Lin, Y.-C.; Deng, F.; Liu, S.-B. New Insights into Keggin-Type 12-Tungstophosphoric Acid from 31P MAS NMR Analysis of Absorbed Trimethylphosphine Oxide and DFT Calculations Chem. - Asian J. 2011, 6, 137– 148 DOI: 10.1002/asia.201000572Google Scholar41New Insights into Keggin-Type 12-Tungstophosphoric Acid from 31P MAS NMR Analysis of Absorbed Trimethylphosphine Oxide and DFT CalculationsHuang, Shing-Jong; Yang, Chih-Yi; Zheng, Anmin; Feng, Ningdong; Yu, Ningya; Wu, Pei-Hao; Chang, Yu-Chi; Lin, Ying-Chih; Deng, Feng; Liu, Shang-BinChemistry - An Asian Journal (2011), 6 (1), 137-148CODEN: CAAJBI; ISSN:1861-4728. (Wiley-VCH Verlag GmbH & Co. KGaA)The acid and transport properties of the anhyd. Keggin-type 12-tungstophosphoric acid (H3PW12O40; HPW) have been studied by solid-state 31P magic-angle spinning NMR of absorbed trimethylphosphine oxide (TMPO) in conjunction with DFT calcns. Accordingly, 31P NMR resonances arising from various protonated complexes, such as TMPOH+ and (TMPO)2H+ adducts, could be unambiguously identified. It was found that thermal pretreatment of the sample at elevated temps. (≥423 K) is a prerequisite for ensuring complete penetration of the TMPO guest probe mol. into HPW particles. Transport of the TMPO absorbate into the matrix of the HPW adsorbent was found to invoke a desorption/absorption process assocd. with the (TMPO)2H+ adducts. Consequently, three types of protonic acid sites with distinct superacid strengths, which correspond to 31P chem. shifts of 92.1, 89.4, and 87.7 ppm, were obsd. for HPW samples loaded with less than three mols. of TMPO per Keggin unit. Together with detailed DFT calcns., these results support the scenario that the TMPOH+ complexes are assocd. with protons located at three different terminal oxygen (Od) sites of the PW12O403- polyanions. Upon increasing the TMPO loading to >3.0 mols. per Keggin unit, abrupt decreases in acid strength and the corresponding structural variations were attributed to the change in secondary structure of the pseudoliquid phase of HPW in the presence of excessive guest absorbate.
Cited By
Smart citations by scite.ai include citation statements extracted from the full text of the citing article. The number of the statements may be higher than the number of citations provided by ACS Publications if one paper cites another multiple times or lower if scite has not yet processed some of the citing articles.
This article is cited by 26 publications.
- Ankur Agarwal, Jeffrey D. Rimer, Jeremy C. Palmer. Water Structure and Dynamics Near the Surfaces of Silicalite-1. The Journal of Physical Chemistry C 2023, 127
(49)
, 23887-23896. https://doi.org/10.1021/acs.jpcc.3c05723
- Michael Dyballa. Solid-State NMR Probe Molecules for Catalysts and Adsorbents: Concepts, Quantification, Accessibility, and Spatial Distribution. Energy & Fuels 2023, 37
(23)
, 18517-18559. https://doi.org/10.1021/acs.energyfuels.3c03815
- Céline Chizallet, Christophe Bouchy, Kim Larmier, Gerhard Pirngruber. Molecular Views on Mechanisms of Brønsted Acid-Catalyzed Reactions in Zeolites. Chemical Reviews 2023, 123
(9)
, 6107-6196. https://doi.org/10.1021/acs.chemrev.2c00896
- Putla Sudarsanam, Navneet Kumar Gupta, Baithy Mallesham, Nittan Singh, Pavan Narayan Kalbande, Benjaram M. Reddy, Bert F. Sels. Supported MoOx and WOx Solid Acids for Biomass Valorization: Interplay of Coordination Chemistry, Acidity, and Catalysis. ACS Catalysis 2021, 11
(21)
, 13603-13648. https://doi.org/10.1021/acscatal.1c03326
- Carlos Bornes, Michael Fischer, Jeffrey A. Amelse, Carlos F. G. C. Geraldes, João Rocha, Luís Mafra. What Is Being Measured with P-Bearing NMR Probe Molecules Adsorbed on Zeolites?. Journal of the American Chemical Society 2021, 143
(34)
, 13616-13623. https://doi.org/10.1021/jacs.1c05014
- Gaurav Kumar, Limin Ren, Yutong Pang, Xinyu Li, Han Chen, Jason Gulbinski, Paul J. Dauenhauer, Michael Tsapatsis, Omar A. Abdelrahman. Acid Sites of Phosphorus-Modified Zeosils. ACS Catalysis 2021, 11
(15)
, 9933-9948. https://doi.org/10.1021/acscatal.1c01588
- Wenjie Yang, Zichun Wang, Jun Huang, Yijiao Jiang. Qualitative and Quantitative Analysis of Acid Properties for Solid Acids by Solid-State Nuclear Magnetic Resonance Spectroscopy. The Journal of Physical Chemistry C 2021, 125
(19)
, 10179-10197. https://doi.org/10.1021/acs.jpcc.1c01887
- Yongxiang Wang, Shaohui Xin, Yueying Chu, Jun Xu, Guodong Qi, Qiang Wang, Qinghua Xia, Feng Deng. Influence of Trimethylphosphine Oxide Loading on the Measurement of Zeolite Acidity by Solid-State NMR Spectroscopy. The Journal of Physical Chemistry C 2021, 125
(17)
, 9497-9506. https://doi.org/10.1021/acs.jpcc.1c01789
- Mariya Shamzhy, Barbara Gil, Maksym Opanasenko, Wieslaw J. Roth, Jiří Čejka. MWW and MFI Frameworks as Model Layered Zeolites: Structures, Transformations, Properties, and Activity. ACS Catalysis 2021, 11
(4)
, 2366-2396. https://doi.org/10.1021/acscatal.0c05332
- Céline Chizallet. Toward the Atomic Scale Simulation of Intricate Acidic Aluminosilicate Catalysts. ACS Catalysis 2020, 10
(10)
, 5579-5601. https://doi.org/10.1021/acscatal.0c01136
- Carlos Hernandez-Tamargo, Alberto Roldan, Nora H. de Leeuw. Tautomerization of Phenol at the External Lewis Acid Sites of Scandium-, Iron- and Gallium-Substituted Zeolite MFI. The Journal of Physical Chemistry C 2019, 123
(13)
, 7604-7614. https://doi.org/10.1021/acs.jpcc.8b02455
- Alexander Hoffman, Mykela DeLuca, David Hibbitts. Restructuring of MFI Framework Zeolite Models and Their Associated Artifacts in Density Functional Theory Calculations. The Journal of Physical Chemistry C 2019, 123
(11)
, 6572-6585. https://doi.org/10.1021/acs.jpcc.8b12230
- Anmin Zheng, Shang-Bin Liu, and Feng Deng . 31P NMR Chemical Shifts of Phosphorus Probes as Reliable and Practical Acidity Scales for Solid and Liquid Catalysts. Chemical Reviews 2017, 117
(19)
, 12475-12531. https://doi.org/10.1021/acs.chemrev.7b00289
- Juncong Yuan, Yaqi Dong, ZePing Wang, Xuliang Deng, Ruiying Li, Jiaxu Wang, Yuchang Zhu, Zhe Ma, Yibin Liu, De Chen, Chaohe Yang, Xiang Feng. Tailoring the surface islands to modulate the diffusion of guest molecules over the TS-1 zeolites for enhanced selectivity in the 1-hexene epoxidation. Chemical Engineering Journal 2025, 503 , 158279. https://doi.org/10.1016/j.cej.2024.158279
- Carlos Bornes, Carlos F.G.C. Geraldes, João Rocha, Luís Mafra. Investigating trimethylphosphine oxide interactions with Brønsted and Lewis acid sites in zeolites: A comparative solid-state NMR study of wet-phase and gas-phase adsorption techniques. Microporous and Mesoporous Materials 2023, 360 , 112666. https://doi.org/10.1016/j.micromeso.2023.112666
- Pui Ching Lan, Yin Zhang, Weijie Zhang, Xueying Ge, Shengqian Ma. Mimicking Enzymatic Non‐Covalent Interactions with Functionalized Covalent Organic Frameworks for Improved Adsorption and Hydrolysis of Cellobiose. Macromolecular Rapid Communications 2023, 44
(11)
https://doi.org/10.1002/marc.202200724
- Pui Ching Lan, Shengqian Ma. Cleaner Futures: Covalent Organic Frameworks for Sustainable Degradation of Lignocellulosic Materials. 2023https://doi.org/10.12794/metadc2137620
- Zheng Li, Michael Benz, Carolin Rieg, Daniel Dittmann, Ann-Katrin Beurer, Dorothea Häussermann, Bjørnar Arstad, Michael Dyballa. The alumination mechanism of porous silica materials and properties of derived ion exchangers and acid catalysts. Materials Chemistry Frontiers 2021, 5
(11)
, 4254-4271. https://doi.org/10.1039/D1QM00282A
- Esun Selvam, Rajesh K. Parsapur, Carlos E. Hernandez-Tamargo, Nora H. de Leeuw, Parasuraman Selvam. Nanostructured zeolite with brain-coral morphology and tailored acidity: a self-organized hierarchical porous material with MFI topology. CrystEngComm 2020, 22
(38)
, 6275-6286. https://doi.org/10.1039/D0CE00989J
- Xianfeng Yi, Hui-Hsin Ko, Feng Deng, Shang-Bin Liu, Anmin Zheng. Solid-state 31P NMR mapping of active centers and relevant spatial correlations in solid acid catalysts. Nature Protocols 2020, 15
(10)
, 3527-3555. https://doi.org/10.1038/s41596-020-0385-6
- Carlos Bornes, Mariana Sardo, Zhi Lin, Jeffrey Amelse, Auguste Fernandes, Maria Filipa Ribeiro, Carlos Geraldes, João Rocha, Luís Mafra. 1
H–
31
P HETCOR NMR elucidates the nature of acid sites in zeolite HZSM-5 probed with trimethylphosphine oxide. Chemical Communications 2019, 55
(84)
, 12635-12638. https://doi.org/10.1039/C9CC06763A
- Christopher J. Heard, Jiří Čejka, Maksym Opanasenko, Petr Nachtigall, Gabriele Centi, Siglinda Perathoner. 2D Oxide Nanomaterials to Address the Energy Transition and Catalysis. Advanced Materials 2019, 31
(3)
https://doi.org/10.1002/adma.201801712
- Wei‐Che Lin, Lin Ye, Simson Wu, Ben Lo, Yung‐Kang Peng, Pu Zhao, Ian McPherson, Shik Chi Edman Tsang. Zinc‐Incorporated Microporous Molecular Sieve for Mild Catalytic Hydrolysis of γ‐Valerolactone: A New Selective Route for Biomass Conversion. ChemSusChem 2018, 11
(24)
, 4214-4218. https://doi.org/10.1002/cssc.201802079
- Carlos E. Hernandez-Tamargo, Alberto Roldan, Phuti E. Ngoepe, Nora H. de Leeuw. Periodic modeling of zeolite Ti-LTA. The Journal of Chemical Physics 2017, 147
(7)
https://doi.org/10.1063/1.4998296
- Leslie Reguera, Noeldris L. López, Joelis Rodríguez‐Hernández, Marlene González, Carlos E. Hernandez‐Tamargo, David Santos‐Carballal, Nora H. de Leeuw, Edilso Reguera. Synthesis, Crystal Structures, and Properties of Zeolite‐Like T
3
(H
3
O)
2
[M(CN)
6
]
2
·
u
H
2
O (T = Co, Zn; M = Ru, Os). European Journal of Inorganic Chemistry 2017, 2017
(23)
, 2980-2989. https://doi.org/10.1002/ejic.201700278
- , E. M. Demianenko, A. G. Grebenyuk, , V. V. Lobanov, , O. S. Kukolevska, , I. I. Gerashchenko, , M. I. Terets, . Quantum chemical modeling of adsorption complexes of fragment of poly(2 hydroxyethyl methacrylate) on silica surface. Surface 2016, 8(23) , 78-84. https://doi.org/10.15407/Surface.2016.08.078
Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.
Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.
The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.
Recommended Articles
Abstract
Figure 1
Figure 1. Representation of the BA sites at the internal (bottom, left panel) and at the external (bottom, right panel) surfaces. One of the two pentasil layers that form the slab is identified by a black-line rectangle (top, right panel). The Al-substituted T2-site (light blue ball) with the proton (white ball) at the O1 position (red ball) are shown. The rest of the O atoms and silanol OH groups were deleted for a better view; Si atoms are represented by orange sticks. A molecule of trimethylphosphine oxide is shown in the bottom-right corner; H is represented in white, C in gray, P in brown, and O in red.
Figure 2
Figure 2. Correlation of the O(P)–H distances and the 31P chemical shifts using Zheng et al. data (black squares linked with black lines). (11) The experimental classification of the acid sites of the zeolite MFI according to the TMPO 31P chemical shift is indicated by vertical dashed blue lines. (10, 12) The experimental classification is used to extrapolate the expected O(P)–H distances from its interception with the theoretical curve (indicated by red circles); the corresponding distance values are written above the horizontal dashed black lines. The spectrum of O(P)–H distances is divided into three zones taking the middle points between the extrapolated O(P)–H distances. These zones are shaded alternately in light gray and white, corresponding to (top) weak acids, (center) strong acids, and (bottom) very strong acids; the limits of each range are indicated by the red numbers at the left-hand side of the graph.
Figure 3
Figure 3. Representation of the interaction of a single TMPO molecule with (a, b) the internal and (c, d) the external BA sites after local optimization; H in white, C in gray, P in brown, O in red, Al in light blue, and Si represented by orange sticks. All the framework O atoms (except the protonated one) and silanol OH groups were deleted for an enhanced view. Related structural values are presented in Table 1.
Figure 4
Figure 5
Figure 5. Representation of the adsorption of two TMPO molecules on (a) the internal and (b) the external BA sites after local optimization. H in white, C in gray, P in brown, O in red, Al in light blue, and Si represented by orange sticks. All the framework O atoms (except the protonated one) and silanol OH groups were deleted for an enhanced view. Related structural values are presented in Table 3.
Figure 6
Figure 7
Figure 7. Representation of the adsorption of three TMPO molecules on (a) the internal and (b) the external BA sites after local optimization. H in white, C in gray, P in brown, O in red, Al in light blue, and Si represented by orange sticks. All the framework O atoms (except the protonated one) and silanol OH groups were deleted for an enhanced view. The interaction between one of the nonprotonated TMPO and the Al atom is represented by a stick connecting O(P2) to the Al atom. Related structural values are presented in Table 5.
Figure 8
Figure 8. Charge difference isosurfaces with values of 0.005 bohr–3 calculated from eq 1. (a) Internal BA site; (b) external BA site. The structures correspond to those shown in Figure 7. H in white, C in gray, P in brown, O in red, Al in light blue, and Si represented by orange sticks. All the framework O atoms (except the protonated one) and silanol OH groups were deleted for an enhanced view.
Figure 9
Figure 10
Figure 10. Representation of (a, c) TMPOH+ and (b, d) (TMPO)2H+ at more than 5 Å away from the BA site after local optimization. (a, b) Internal surface; (c, d) external surface. H in white, C in gray, P in brown, O in red, and Si represented by orange sticks. All the framework O atoms (except the protonated one) and silanol OH groups were deleted for an enhanced view. Related structural values are presented in Table 8.
References
This article references 41 other publications.
- 1Sandoval-Díaz, L.-E.; González-Amaya, J.-A.; Trujillo, C.-A. General Aspects of Zeolite Acidity Characterization Microporous Mesoporous Mater. 2015, 215, 229– 243 DOI: 10.1016/j.micromeso.2015.04.0381General aspects of zeolite acidity characterizationSandoval-Diaz, Luis-Ernesto; Gonzalez-Amaya, Jhon-Alex; Trujillo, Carlos-AlexanderMicroporous and Mesoporous Materials (2015), 215 (), 229-243CODEN: MIMMFJ; ISSN:1387-1811. (Elsevier Inc.)A review; this paper describes the most used techniques in the detn. of the acid property in zeolite materials. Two families of techniques, namely spectrometric (IR, NMR) and adsorption-desorption methods (calorimetry, TPD) are considered. Typical exptl. conditions, schematics of equipment setup, and rules for mol. probe selection are shown. Selected expts. that make use of these methods are briefly discussed, and their most relevant results are presented. This review shows some of the possibilities that can be used to "measure" acidity in zeolites. It is also shown that the interpretation of such results could be deceiving in certain situations, esp. because acidity and catalytic performance are not synonyms and acidity is inherently a relative term.
- 2Vermeiren, W.; Gilson, J.-P. Impact of Zeolites on the Petroleum and Petrochemical Industry Top. Catal. 2009, 52, 1131– 1161 DOI: 10.1007/s11244-009-9271-82Impact of Zeolites on the Petroleum and Petrochemical IndustryVermeiren, W.; Gilson, J.-P.Topics in Catalysis (2009), 52 (9), 1131-1161CODEN: TOCAFI; ISSN:1022-5528. (Springer)A review. The general features of zeolites that led to their widespread use in oil refining and petrochem. are highlighted as well as the details of their impact on selected processes. The anal. of the catalyst market and the position of zeolites therein is a good indication of their strategic importance. Zeolites have brought many disruptive changes to these fields (e.g. FCC). They impacted also these industries in an equally important way, although more subtle, by incremental improvement of processes. The new and vast challenges facing oil refining and petrochem. as well as the managed transition to sustainable environmental benign transport fuel industries and chem. industries will require creative science and technologies. Zeolites offer the basis of many of these technol. solns. provided efficient and balanced cooperations between industry and academia are further developed.
- 3Weitkamp, J. Zeolites and Catalysis Solid State Ionics 2000, 131, 175– 188 DOI: 10.1016/S0167-2738(00)00632-93Zeolites and catalysisWeitkamp, J.Solid State Ionics (2000), 131 (1,2), 175-188CODEN: SSIOD3; ISSN:0167-2738. (Elsevier Science B.V.)This review with 37 refs. covers the fundamentals of zeolite materials science and their application as catalysts. After a brief introduction into their structures, the most important parameters are discussed which allow the prepn. of an almost infinite variety of zeolitic materials tailored for a given catalytic application. Zeolites are solid acids, and the chem. nature, the d., strength and location of the acid sites in zeolites are discussed. Shape-selective catalysis, which is a unique feature of zeolites, is briefly addressed.
- 4Moliner, M.; Martínez, C.; Corma, A. Multipore Zeolites: Synthesis and Catalytic Applications Angew. Chem., Int. Ed. 2015, 54, 3560– 3579 DOI: 10.1002/anie.2014063444Multipore Zeolites: Synthesis and Catalytic ApplicationsMoliner, Manuel; Martinez, Cristina; Corma, AvelinoAngewandte Chemie, International Edition (2015), 54 (12), 3560-3579CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)A review; in the last few years, important efforts have been made to synthesize so-called "multipore" zeolites, which contain channels of different dimensions within the same cryst. structure. This is a very attractive subject, since the presence of pores of different sizes would favor the preferential diffusion of reactants and products through those different channel systems, allowing unique catalytic activities for specific chem. processes. In this Review we describe the most attractive achievements in the rational synthesis of multipore zeolites, contg. small to extra-large pores, and the improvements reported for relevant chem. processes when these multipore zeolites have been used as catalysts.
- 5Čejka, J.; Wichterlová, B. Acid-Catalyzed Synthesis of Mono- and Dialkyl Benzenes Over Zeolites: Active Sites, Zeolite Topology, and Reaction Mechanisms Catal. Rev.: Sci. Eng. 2002, 44, 375– 421 DOI: 10.1081/CR-1200057415Acid-catalyzed synthesis of mono- and dialkyl benzenes over zeolites: active sites, zeolite topology, and reaction mechanismsCejka, Jiri; Wichterlova, BlankaCatalysis Reviews - Science and Engineering (2002), 44 (3), 375-421CODEN: CRSEC9; ISSN:0161-4940. (Marcel Dekker, Inc.)A review of mechanisms of catalytic reactions leading to mono- and dialkyl benzenes in the presence of zeolite catalysts which are an important part of petrochem. prodn. Emphasis is on the effect of the type of acid sites, zeolite structure, and reaction conditions on the activity and selectivity of these complex reactions, and particularly on the individual products with respect to their iso- vs. n- and ortho-, meta-, and para-isomers. First, descriptions and anal. were made of the structure and properties of acid sites and zeolite pore inner structures as applied in the synthesis of alkyl benzenes. Different structural types of medium- and large-pore zeolite mol. sieves were tested in C6H6 and PhMe alkylation with MeCH:CH2 and Me2CHOH to give cumene and cymenes. Secondly, individual reactions leading to the synthesis of PhPr, xylenes, ethyltoluenes, diethylbenzenes, and transformation of trimethylbenzenes to xylenes from the viewpoint of their reaction mechanism, function of the acid sites, and inner pore geometry of zeolites were described.
- 6Derouane, E. G.; Védrine, J. C.; Pinto, R. R.; Borges, P. M.; Costa, L.; Lemos, M. A. N. D. A.; Lemos, F.; Ribeiro, F. R. The Acidity of Zeolites: Concepts, Measurements and Relation to Catalysis: A Review on Experimental and Theoretical Methods for the Study of Zeolite Acidity Catal. Rev.: Sci. Eng. 2013, 55, 454– 515 DOI: 10.1080/01614940.2013.8222666The acidity of zeolites: Concepts, measurements and relation to catalysis: A review on experimental and theoretical methods for the study of zeolite acidityDerouane, E. G.; Vedrine, J. C.; Pinto, R. Ramos; Borges, P. M.; Costa, L.; Lemos, M. A. N. D. A.; Lemos, F.; Ribeiro, F. RamoaCatalysis Reviews: Science and Engineering (2013), 55 (4), 454-515CODEN: CRSEC9; ISSN:0161-4940. (Taylor & Francis, Inc.)A review. Considered are all aspects of acidity (nature of acid sites, strength, d., etc.) in solid catalysts and in zeolites in particular. After reminding the definition of acidity in liq. and solid acids, the authors emphasized acidity characterization by the most used phys. techniques, such as Hammett's indicator titrn., microcalorimetry of adsorbed probe mols. (ammonia, pyridine or other amines for acidity characterization and CO2 or SO2 for basicity characterization), ammonia or any amine thermodesorption, IR spectroscopy of hydroxyl groups and of several probe mols. adsorbed (ammonia, pyridine, piperidine, amines, CO, H2, etc.), MAS-NMR of 27Al, 29Si, 1H elements and of 1H, 13C, 31P, etc. of adsorbed probe mols., and model catalytic reactions. Modeling the way the acid features of zeolites influence the catalytic activity of these catalysts toward acid-catalyzed reactions (relation between ammonia desorption activation energy values and catalytic activities, reaction mechanism, and kinetics) completes the general anal. of acidity and zeolite chem.
- 7Zheng, A.; Liu, S.-B.; Deng, F. Acidity Characterization of Heterogeneous Catalysts by Solid-State NMR Spectroscopy Using Probe Molecules Solid State Nucl. Magn. Reson. 2013, 55–56, 12– 27 DOI: 10.1016/j.ssnmr.2013.09.0017Acidity characterization of heterogeneous catalysts by solid-state NMR spectroscopy using probe moleculesZheng, Anmin; Liu, Shang-Bin; Deng, FengSolid State Nuclear Magnetic Resonance (2013), 55-56 (), 12-27CODEN: SSNRE4; ISSN:0926-2040. (Elsevier)A review; characterization of the surface acidic properties of solid acid catalysts is a key issue in heterogeneous catalysis. Important acid features of solid acids, such as their type (Bronsted vs. Lewis acid), distribution and accessibility (internal vs. external sites), concn. (amt.), and strength of acid sites are crucial factors dictating their reactivity and selectivity. This short review provides information on different solid-state NMR techniques used for acidity characterization of solid acid catalysts. In particular, different approaches using probe mols. contg. a specific nucleus of interest, such as pyridine-d5, 2-13C-acetone, trimethylphosphine, and trimethylphosphine oxide, are compared. Incorporation of valuable information (such as the adsorption structure, deprotonation energy, and NMR parameters) from d. functional theory (DFT) calcns. can yield explicit correlations between the chem. shift of adsorbed probe mols. and the intrinsic acid strength of solid acids. Methods that combine exptl. NMR data with DFT calcns. can therefore provide both qual. and quant. information on acid sites.
- 8Xu, J.; Zheng, A.; Yang, J.; Su, Y.; Wang, J.; Zeng, D.; Zhang, M.; Ye, C.; Deng, F. Acidity of Mesoporous MoOx/ZrO2 and WOx/ZrO2Materials: A Combined Solid-State NMR and Theoretical Calculation Study J. Phys. Chem. B 2006, 110, 10662– 10671 DOI: 10.1021/jp06140878Acidity of Mesoporous MoOx/ZrO2 and WOx/ZrO2 Materials: A Combined Solid-State NMR and Theoretical Calculation StudyXu, Jun; Zheng, Anmin; Yang, Jun; Su, Yongchao; Wang, Jiqing; Zeng, Danlin; Zhang, Mingjin; Ye, Chaohui; Deng, FengJournal of Physical Chemistry B (2006), 110 (22), 10662-10671CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)The acidity of mesoporous MoOx/ZrO2 and WOx/ZrO2 materials was studied in detail by multinuclear solid-state NMR techniques as well as DFT quantum chem. calcns. The 1H MAS NMR expts. clearly revealed two different types of strong Bronsted acid sites on both MoOx/ZrO2 and WOx/ZrO2 mesoporous materials, which were able to prontonate adsorbed pyrine-d5 (resulting in 1H NMR signals at chem. shifts in the range 16-19 ppm) as well as adsorbed trimethylphosphine (giving rise to 31P NMR signal at ∼0 ppm). The 13C NMR of adsorbed 2-13C-acetone indicated that the av. Bronsted acid strength of the two mesoporous materials was stronger than that of zeolite HZSM-5 but still weaker than that of 100% H2SO4, which was in good agreement with theor. predictions. The quantum chem. calcns. revealed the detailed structures of the two distinct types of Bronsted acid sites formed on the mesoporous MoOx/ZrO2 and WOx/ZrO2. The existence of both monomer and oligomer Mo (or W) species contg. a Mo-OH-Zr (or W-OH-Zr) bridging OH group was confirmed with the former having an acid strength close to zeolite HZSM-5, with the latter having an acid strength similar to sulfated zirconia. From the authors' NMR exptl. and theor. calcn. results, a possible mechanism is proposed for the formation of acid sites on these mesoporous materials.
- 9Li, S.; Huang, S.-J.; Shen, W.; Zhang, H.; Fang, H.; Zheng, A.; Liu, S.-B.; Deng, F. Probing the Spatial Proximities among Acid Sites in Dealuminated H-Y Zeolite by Solid-State NMR Spectroscopy J. Phys. Chem. C 2008, 112, 14486– 14494 DOI: 10.1021/jp803494n9Probing the Spatial Proximities among Acid Sites in Dealuminated H-Y Zeolite by Solid-State NMR SpectroscopyLi, Shenhui; Huang, Shing-Jong; Shen, Wanling; Zhang, Hailu; Fang, Hanjun; Zheng, Anmin; Liu, Shang-Bin; Deng, FengJournal of Physical Chemistry C (2008), 112 (37), 14486-14494CODEN: JPCCCK; ISSN:1932-7447. (American Chemical Society)A comprehensive study has been made to probe the spatial proximities among different acid sites in dealuminated H-Y zeolites modified with various degrees of calcination, steam, and acid treatments by using a variety of different solid-state NMR techniques, including multinuclear MAS NMR and two-dimensional 1H double-quantum (DQ) MAS NMR spectroscopy. The effects of dealumination treatments on the nature, concn., and location of extraframework Al species in H-Y zeolites were followed by 1H DQ MAS NMR of hydroxyl protons in conjunction with 1H, 27Al, and 29Si MAS NMR results. It was found that the extraframework AlOH species (Lewis acid sites) are always in close proximity to the bridging AlOHSi hydroxyls (Bronsted acid sites) on the framework of dealuminated H-Y zeolites prepd. by thermal and hydrothermal treatments, indicating the presence of a Bronsted/Lewis acid synergy effect. However, such an effect is absent in acid-treated H-Y zeolites, as also confirmed by 13C CP/MAS NMR of adsorbed 2-13C-acetone.
- 10Zhao, Q.; Chen, W.-H.; Huang, S.-J.; Wu, Y.-C.; Lee, H.-K.; Liu, S.-B. Discernment and Quantification of Internal and External Acid Sites on Zeolites J. Phys. Chem. B 2002, 106, 4462– 4469 DOI: 10.1021/jp015574k10Discernment and Quantification of Internal and External Acid Sites on ZeolitesZhao, Qi; Chen, Wen-Hua; Huang, Shing-Jong; Wu, Yu-Chih; Lee, Huang-Kuei; Liu, Shang-BinJournal of Physical Chemistry B (2002), 106 (17), 4462-4469CODEN: JPCBFK; ISSN:1089-5647. (American Chemical Society)A new methodol. is reported for concurrent qual. and quant. characterization of internal and external acid sites in zeolitic catalysts. H-ZSM-5 zeolites with varied Si/Al ratios have been examd. by solid-state 31P MAS NMR using different adsorbed probe mols., namely trimethylphosphine oxides (TMPO) and tributylphosphine oxide (TBPO), in conjunction with elemental anal. Up to seven distinct 31P resonance peaks at 86, 75, 67, 63, 53, 43, and 30 ppm were identified from the 31P NMR spectra of adsorbed TMPO. The resonance peak at 30 ppm has never been obsd. previously and may be ascribed to mobile TMPO. The peak at 43 ppm is assigned to physisorbed TMPO. The rest of the peaks result from TMPOH+ complexes at Bronsted sites with peaks at higher chem. shifts reflecting acid sites of higher strengths. 31P NMR expts., performed with TMPO and TBPO adsorbed on a mesoporous MCM-41 sample, resp., provide further correlation of the internal and external acid sites. It is concluded that the peaks at 75 and 53 ppm arise exclusively from the internal Bronsted sites, whereas the peaks at 86, 67, and 63 ppm are assocd. with both internal and external acid sites. While the concn. and distribution of internal sites were found to increase with acid strengths, changing the Si/Al ratio of HZSM-5 has nearly no effect on the strength of the external acid sites.
- 11Zheng, A.; Zhang, H.; Lu, X.; Liu, S.-B.; Deng, F. Theoretical Predictions of 31P NMR Chemical Shift Threshold of Trimethylphosphine Oxide Absorbed on Solid Acid Catalysts J. Phys. Chem. B 2008, 112, 4496– 4505 DOI: 10.1021/jp709739v11Theoretical Predictions of 31P NMR Chemical Shift Threshold of Trimethylphosphine Oxide Absorbed on Solid Acid CatalystsZheng, Anmin; Zhang, Hailu; Lu, Xin; Liu, Shang-Bin; Deng, FengJournal of Physical Chemistry B (2008), 112 (15), 4496-4505CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)The 31P NMR chem. shifts of adsorbed trimethylphosphine oxide (TMPO) and the configurations of the corresponding TMPOH+ complexes on Bronsted acid sites with varying acid strengths in modeled zeolites have been predicted theor. by means of d. functional theory (DFT) quantum chem. calcns. The configuration of each TMPOH+ complex was optimized at the PW91/DNP level based on an 8T cluster model, whereas the 31P chem. shifts were calcd. with the gauge including AO (GIAO) approach at both the HF/TZVP and MP2/TZVP levels. A linear correlation between the 31P chem. shift of adsorbed TMPO and the proton affinity of the solid acids was obsd., and a threshold for superacidity (86 ppm) was detd. This threshold for superacidity was also confirmed by comparative investigations on other superacid systems, such as carborane acid and heteropolyoxometalate H3PW12O40. In conjunction with the strong correlation between the MP2 and the HF 31P isotropic shifts, the 8T cluster model was extended to more sophisticated models (up to 72T) that are not readily tractable at the GIAO-MP2 level, and a 31P chem. shift of 86 ppm was detd. for TMPO adsorbed on zeolite H-ZSM-5, which is in good agreement with the NMR exptl. data.
- 12Seo, Y.; Cho, K.; Jung, Y.; Ryoo, R. Characterization of the Surface Acidity of MFI Zeolite Nanosheets by 31 P NMR of Adsorbed Phosphine Oxides and Catalytic Cracking of Decalin ACS Catal. 2013, 3, 713– 720 DOI: 10.1021/cs300824e12Characterization of the Surface Acidity of MFI Zeolite Nanosheets by 31P NMR of Adsorbed Phosphine Oxides and Catalytic Cracking of DecalinSeo, Yongbeom; Cho, Kanghee; Jung, Younjae; Ryoo, RyongACS Catalysis (2013), 3 (4), 713-720CODEN: ACCACS; ISSN:2155-5435. (American Chemical Society)MFI zeolite nanosheets tailored to 2.5-nm thickness were synthesized using a surfactant-type zeolite structure-directing agent, [C22H45-N+(CH3)2-C6H12-N+(CH3)2-C6H13](Br-)2. The zeolite nanosheets possessed Bronsted acid sites on their external surfaces as well as in the internal micropore walls. The acid strength and concn. was characterized by the 31P NMR signals of the adsorbed trimethylphosphine oxide and tributylphosphine oxide. The 31P NMR investigation identified three types of Bronsted acid sites with different strengths on external surfaces; there were four types inside the micropores. A linear correlation has been established between the no. of the external strongest acid sites and the catalytic activity in decalin cracking for the MFI zeolite catalysts investigated in this work.
- 13Ni, Y.; Sun, A.; Wu, X.; Hai, G.; Hu, J.; Li, T.; Li, G. The Preparation of Nano-Sized H[Zn,Al]ZSM-5 Zeolite and Its Application in the Aromatization of Methanol Microporous Mesoporous Mater. 2011, 143, 435– 442 DOI: 10.1016/j.micromeso.2011.03.029There is no corresponding record for this reference.
- 14Costa, C.; Dzikh, I. P.; Lopes, J. M.; Lemos, F.; Ribeiro, F. R. Activity–acidity Relationship in Zeolite ZSM-5. Application of Brönsted-Type Equations J. Mol. Catal. A: Chem. 2000, 154, 193– 201 DOI: 10.1016/S1381-1169(99)00374-X14Activity-acidity relationship in zeolite ZSM-5. Application of Bronsted-type equationsCosta, C.; Dzikh, I. P.; Lopes, J. M.; Lemos, F.; Ribeiro, F. R.Journal of Molecular Catalysis A: Chemical (2000), 154 (1-2), 193-201CODEN: JMCCF2; ISSN:1381-1169. (Elsevier Science B.V.)In this paper the relation between activity and acidity in a variety of ZSM-5 zeolite catalysts, with different Si/Al ratios and different protonic content, is analyzed and a quant. correlation is obtained. The acid site strength distribution was estd. using temp.-programmed desorption (TPD) of ammonia by applying a digital deconvolution method to the curves. These data were then correlated with exptl. catalytic activity data for the same catalysts towards n-heptane cracking reaction, by means of a Bronsted-type equation similar to the ones used for homogeneous acid catalysis and already used for other zeolites. It can be noticed that the same types of equation that are used for homogeneous acid catalysis also hold for heterogeneous acid catalysis and that the activation energy for ammonia desorption can be used as acid-strength scale for the purpose of correlation with catalytic activity.
- 15Kwok, T. J.; Jayasuriya, K.; Damavarapu, R.; Brodman, B. W. Application of H-ZSM-5 Zeolite for Regioselective Mononitration of Toluene J. Org. Chem. 1994, 59, 4939– 4942 DOI: 10.1021/jo00096a042There is no corresponding record for this reference.
- 16Wang, Y.; Jian, G.; Peng, Z.; Hu, J.; Wang, X.; Duan, W.; Liu, B. Preparation and Application of Ba/ZSM-5 Zeolite for Reaction of Methyl Vinyl Ether and Methanol Catal. Commun. 2015, 66, 34– 37 DOI: 10.1016/j.catcom.2015.03.01216Preparation and application of Ba/ZSM-5 zeolite for reaction of methyl vinyl ether and methanolWang, Yang; Jian, Guixing; Peng, Zhihong; Hu, Jiaji; Wang, Xin; Duan, Wubiao; Liu, BoCatalysis Communications (2015), 66 (), 34-37CODEN: CCAOAC; ISSN:1566-7367. (Elsevier B.V.)The ZSM-5 zeolite is widely used to catalyze the reactions of methanol to olefins. Herein, we have prepd. the H-ZSM-5 doped with barium (Ba/ZSM-5) using incipient wetness impregnation method. The Ba modified catalysts were used to catalyze a new reaction of methanol with Me vinyl ether to improve the selectivity of ethylene and propylene (C=2 + C=3). The reaction catalyzed by Ba doped H-ZSM-5 shows higher propylene selectivity over H-ZSM-5. The reaction mechanism is discussed.
- 17Zheng, A.; Han, B.; Li, B.; Liu, S.-B.; Deng, F. Enhancement of Brønsted Acidity in Zeolitic Catalysts due to an Intermolecular Solvent Effect in Confined Micropores Chem. Commun. 2012, 48, 6936– 6938 DOI: 10.1039/c2cc32498a17Enhancement of Bronsted acidity in zeolitic catalysts due to an intermolecular solvent effect in confined microporesZheng, Anmin; Han, Bing; Li, Bojie; Liu, Shang-Bin; Deng, FengChemical Communications (Cambridge, United Kingdom) (2012), 48 (55), 6936-6938CODEN: CHCOFS; ISSN:1359-7345. (Royal Society of Chemistry)Solid-state NMR and DFT calcn. studies certified the presence of an intermol. solvent effect for mols. confined in microporous zeolite, leading to a notable increase in Bronsted acidity of the solid acid catalyst.
- 18Kresse, G.; Hafner, J. Ab Initio Molecular Dynamics for Liquid Metals Phys. Rev. B: Condens. Matter Mater. Phys. 1993, 47, 558– 561 DOI: 10.1103/PhysRevB.47.55818Ab initio molecular dynamics of liquid metalsKresse, G.; Hafner, J.Physical Review B: Condensed Matter and Materials Physics (1993), 47 (1), 558-61CODEN: PRBMDO; ISSN:0163-1829.The authors present ab initio quantum-mech. mol.-dynamics calcns. based on the calcn. of the electronic ground state and of the Hellmann-Feynman forces in the local-d. approxn. at each mol.-dynamics step. This is possible using conjugate-gradient techniques for energy minimization, and predicting the wave functions for new ionic positions using sub-space alignment. This approach avoids the instabilities inherent in quantum-mech. mol.-dynamics calcns. for metals based on the use of a factitious Newtonian dynamics for the electronic degrees of freedom. This method gives perfect control of the adiabaticity and allows one to perform simulations over several picoseconds.
- 19Kresse, G.; Hafner, J. Ab Initio Molecular-Dynamics Simulation of the Liquid-Metal–amorphous-Semiconductor Transition in Germanium Phys. Rev. B: Condens. Matter Mater. Phys. 1994, 49, 14251– 14269 DOI: 10.1103/PhysRevB.49.1425119Ab initio molecular-dynamics simulation of the liquid-metal-amorphous-semiconductor transition in germaniumKresse, G.; Hafner, J.Physical Review B: Condensed Matter and Materials Physics (1994), 49 (20), 14251-69CODEN: PRBMDO; ISSN:0163-1829.The authors present ab initio quantum-mech. mol.-dynamics simulations of the liq.-metal-amorphous-semiconductor transition in Ge. The simulations are based on (a) finite-temp. d.-functional theory of the 1-electron states, (b) exact energy minimization and hence calcn. of the exact Hellmann-Feynman forces after each mol.-dynamics step using preconditioned conjugate-gradient techniques, (c) accurate nonlocal pseudopotentials, and (d) Nose' dynamics for generating a canonical ensemble. This method gives perfect control of the adiabaticity of the electron-ion ensemble and allows the authors to perform simulations over >30 ps. The computer-generated ensemble describes the structural, dynamic, and electronic properties of liq. and amorphous Ge in very good agreement with expt.. The simulation allows the authors to study in detail the changes in the structure-property relation through the metal-semiconductor transition. The authors report a detailed anal. of the local structural properties and their changes induced by an annealing process. The geometrical, bounding, and spectral properties of defects in the disordered tetrahedral network are studied and compared with expt.
- 20Kresse, G.; Furthmüller, J. Efficiency of Ab-Initio Total Energy Calculations for Metals and Semiconductors Using a Plane-Wave Basis Set Comput. Mater. Sci. 1996, 6, 15– 50 DOI: 10.1016/0927-0256(96)00008-020Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis setKresse, G.; Furthmuller, J.Computational Materials Science (1996), 6 (1), 15-50CODEN: CMMSEM; ISSN:0927-0256. (Elsevier)The authors present a detailed description and comparison of algorithms for performing ab-initio quantum-mech. calcns. using pseudopotentials and a plane-wave basis set. The authors will discuss: (a) partial occupancies within the framework of the linear tetrahedron method and the finite temp. d.-functional theory, (b) iterative methods for the diagonalization of the Kohn-Sham Hamiltonian and a discussion of an efficient iterative method based on the ideas of Pulay's residual minimization, which is close to an order N2atoms scaling even for relatively large systems, (c) efficient Broyden-like and Pulay-like mixing methods for the charge d. including a new special preconditioning optimized for a plane-wave basis set, (d) conjugate gradient methods for minimizing the electronic free energy with respect to all degrees of freedom simultaneously. The authors have implemented these algorithms within a powerful package called VAMP (Vienna ab-initio mol.-dynamics package). The program and the techniques have been used successfully for a large no. of different systems (liq. and amorphous semiconductors, liq. simple and transition metals, metallic and semi-conducting surfaces, phonons in simple metals, transition metals and semiconductors) and turned out to be very reliable.
- 21Kresse, G.; Furthmüller, J. Efficient Iterative Schemes for Ab Initio Total-Energy Calculations Using a Plane-Wave Basis Set Phys. Rev. B: Condens. Matter Mater. Phys. 1996, 54, 11169– 11186 DOI: 10.1103/PhysRevB.54.1116921Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis setKresse, G.; Furthmueller, J.Physical Review B: Condensed Matter (1996), 54 (16), 11169-11186CODEN: PRBMDO; ISSN:0163-1829. (American Physical Society)The authors present an efficient scheme for calcg. the Kohn-Sham ground state of metallic systems using pseudopotentials and a plane-wave basis set. In the first part the application of Pulay's DIIS method (direct inversion in the iterative subspace) to the iterative diagonalization of large matrixes will be discussed. This approach is stable, reliable, and minimizes the no. of order Natoms3 operations. In the second part, we will discuss an efficient mixing scheme also based on Pulay's scheme. A special "metric" and a special "preconditioning" optimized for a plane-wave basis set will be introduced. Scaling of the method will be discussed in detail for non-self-consistent and self-consistent calcns. It will be shown that the no. of iterations required to obtain a specific precision is almost independent of the system size. Altogether an order Natoms2 scaling is found for systems contg. up to 1000 electrons. If we take into account that the no. of k points can be decreased linearly with the system size, the overall scaling can approach Natoms. They have implemented these algorithms within a powerful package called VASP (Vienna ab initio simulation package). The program and the techniques have been used successfully for a large no. of different systems (liq. and amorphous semiconductors, liq. simple and transition metals, metallic and semiconducting surfaces, phonons in simple metals, transition metals, and semiconductors) and turned out to be very reliable.
- 22Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple Phys. Rev. Lett. 1996, 77, 3865– 3868 DOI: 10.1103/PhysRevLett.77.386522Generalized gradient approximation made simplePerdew, John P.; Burke, Kieron; Ernzerhof, MatthiasPhysical Review Letters (1996), 77 (18), 3865-3868CODEN: PRLTAO; ISSN:0031-9007. (American Physical Society)Generalized gradient approxns. (GGA's) for the exchange-correlation energy improve upon the local spin d. (LSD) description of atoms, mols., and solids. We present a simple derivation of a simple GGA, in which all parameters (other than those in LSD) are fundamental consts. Only general features of the detailed construction underlying the Perdew-Wang 1991 (PW91) GGA are invoked. Improvements over PW91 include an accurate description of the linear response of the uniform electron gas, correct behavior under uniform scaling, and a smoother potential.
- 23Grimme, S. Semiempirical GGA-Type Density Functional Constructed with a Long-Range Dispersion Correction J. Comput. Chem. 2006, 27, 1787– 1799 DOI: 10.1002/jcc.2049523Semiempirical GGA-type density functional constructed with a long-range dispersion correctionGrimme, StefanJournal of Computational Chemistry (2006), 27 (15), 1787-1799CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)A new d. functional (DF) of the generalized gradient approxn. (GGA) type for general chem. applications termed B97-D is proposed. It is based on Becke's power-series ansatz from 1997 and is explicitly parameterized by including damped atom-pairwise dispersion corrections of the form C6·R-6. A general computational scheme for the parameters used in this correction has been established and parameters for elements up to xenon and a scaling factor for the dispersion part for several common d. functionals (BLYP, PBE, TPSS, B3LYP) are reported. The new functional is tested in comparison with other GGAs and the B3LYP hybrid functional on std. thermochem. benchmark sets, for 40 noncovalently bound complexes, including large stacked arom. mols. and group II element clusters, and for the computation of mol. geometries. Further cross-validation tests were performed for organometallic reactions and other difficult problems for std. functionals. In summary, it is found that B97-D belongs to one of the most accurate general purpose GGAs, reaching, for example for the G97/2 set of heat of formations, a mean abs. deviation of only 3.8 kcal mol-1. The performance for noncovalently bound systems including many pure van der Waals complexes is exceptionally good, reaching on the av. CCSD(T) accuracy. The basic strategy in the development to restrict the d. functional description to shorter electron correlation lengths scales and to describe situations with medium to large interat. distances by damped C6·R-6 terms seems to be very successful, as demonstrated for some notoriously difficult reactions. As an example, for the isomerization of larger branched to linear alkanes, B97-D is the only DF available that yields the right sign for the energy difference. From a practical point of view, the new functional seems to be quite robust and it is thus suggested as an efficient and accurate quantum chem. method for large systems where dispersion forces are of general importance.
- 24Hernandez-Tamargo, C. E.; Roldan, A.; de Leeuw, N. H. A Density Functional Theory Study of the Structure of Pure-Silica and Aluminium-Substituted MFI Nanosheets J. Solid State Chem. 2016, 237, 192– 203 DOI: 10.1016/j.jssc.2016.02.00624A density functional theory study of the structure of pure-silica and aluminium-substituted MFI nanosheetsHernandez-Tamargo, Carlos E.; Roldan, Alberto; de Leeuw, Nora H.Journal of Solid State Chemistry (2016), 237 (), 192-203CODEN: JSSCBI; ISSN:0022-4596. (Elsevier B.V.)The authors present a theor. study of the structural features of silica and aluminum-substituted MFI nanosheets. The authors have analyzed the effects of aluminum substitution on the vibrational properties of silanols as well as the features of protons as counter-ions. The formation of the two-dimensional system did not lead to appreciable distortions within the framework. Moreover, the effects on the structure due to the aluminum dopants were the same in both the bulk and the slab. The principal differences were related to the silanol groups that form hydrogen-bonds with neighboring aluminum-substituted silanols, whereas intra-framework hydrogen-bonds increase the stability of aluminum-substituted silanols toward dehydration. Thus, the authors have complemented previous exptl. and theor. studies, showing the lamellar MFI zeolite to be a very stable material of high crystallinity regardless of its very thin structure.
- 25Blöchl, P. E. Projector Augmented-Wave Method Phys. Rev. B: Condens. Matter Mater. Phys. 1994, 50, 17953– 17979 DOI: 10.1103/PhysRevB.50.1795325Projector augmented-wave methodBlochlPhysical review. B, Condensed matter (1994), 50 (24), 17953-17979 ISSN:0163-1829.There is no expanded citation for this reference.
- 26Kresse, G.; Joubert, D. From Ultrasoft Pseudopotentials to the Projector Augmented-Wave Method Phys. Rev. B: Condens. Matter Mater. Phys. 1999, 59, 1758– 1775 DOI: 10.1103/PhysRevB.59.175826From ultrasoft pseudopotentials to the projector augmented-wave methodKresse, G.; Joubert, D.Physical Review B: Condensed Matter and Materials Physics (1999), 59 (3), 1758-1775CODEN: PRBMDO; ISSN:0163-1829. (American Physical Society)The formal relationship between ultrasoft (US) Vanderbilt-type pseudopotentials and Blochl's projector augmented wave (PAW) method is derived. The total energy functional for US pseudopotentials can be obtained by linearization of two terms in a slightly modified PAW total energy functional. The Hamilton operator, the forces, and the stress tensor are derived for this modified PAW functional. A simple way to implement the PAW method in existing plane-wave codes supporting US pseudopotentials is pointed out. In addn., crit. tests are presented to compare the accuracy and efficiency of the PAW and the US pseudopotential method with relaxed-core all-electron methods. These tests include small mols. (H2, H2O, Li2, N2, F2, BF3, SiF4) and several bulk systems (diamond, Si, V, Li, Ca, CaF2, Fe, Co, Ni). Particular attention is paid to the bulk properties and magnetic energies of Fe, Co, and Ni.
- 27Ho, K.-M.; Fu, C. L.; Harmon, B. N.; Weber, W.; Hamann, D. R. Vibrational Frequencies and Structural Properties of Transition Metals via Total-Energy Calculations Phys. Rev. Lett. 1982, 49, 673– 676 DOI: 10.1103/PhysRevLett.49.67327Vibrational frequencies and structural properties of transition metals via total-energy calculationsHo, K. M.; Fu, C. L.; Harmon, B. N.; Weber, W.; Hamann, D. R.Physical Review Letters (1982), 49 (9), 673-6CODEN: PRLTAO; ISSN:0031-9007.The vibrational frequencies of selected normal modes can be obtained entirely from 1st principles with use of frozen phonon calcns. which involve the precise evaluation of crystal total energy as a function of lattice displacement. The calcns. allow a detailed anal. of the microscopic mechanism causing phonon anomalies and soft-mode phase transitions. Calcns. for Zr, Nb, and Mo were made using both tight-binding and pseudopotential methods.
- 28Fu, C.-L.; Ho, K.-M. First-Principles Calculation of the Equilibrium Ground-State Properties of Transition Metals: Applications to Nb and Mo Phys. Rev. B: Condens. Matter Mater. Phys. 1983, 28, 5480– 5486 DOI: 10.1103/PhysRevB.28.548028First-principles calculation of the equilibrium ground-state properties of transition metals: Application to niobium and molybdenumFu, C. L.; Ho, K. M.Physical Review B: Condensed Matter and Materials Physics (1983), 28 (10), 5480-6CODEN: PRBMDO; ISSN:0163-1829.A self-consistent pseudopotential method was used to calc. the equil. ground-state properties of Mo and Nb. The calcd. equil. lattice consts., cohesive energies, and bulk moduli agree with exptl. data.
- 29Quartieri, S.; Arletti, R.; Vezzalini, G.; Di Renzo, F.; Dmitriev, V. Elastic Behavior of MFI-Type Zeolites: 3 – Compressibility of Silicalite and Mutinaite J. Solid State Chem. 2012, 191, 201– 212 DOI: 10.1016/j.jssc.2012.03.03929Elastic behavior of MFI-type zeolites: 3. Compressibility of silicalite and mutinaiteQuartieri, Simona; Arletti, Rossella; Vezzalini, Giovanna; Di Renzo, Francesco; Dmitriev, VladimirJournal of Solid State Chemistry (2012), 191 (), 201-212CODEN: JSSCBI; ISSN:0022-4596. (Elsevier B.V.)We report the results of an in-situ synchrotron X-ray powder diffraction study, performed using silicone oil as "non-penetrating" pressure transmitting medium, of the elastic behavior of three zeolites with MFI-type framework: the natural zeolite mutinaite and two silicalites (labeled A and B) synthesized under different conditions. In mutinaite, no symmetry change is obsd. as a function of pressure; however, a phase transition from monoclinic (P21/n) to orthorhombic (Pnma) symmetry occurs at about 1.0 GPa in the silicalite samples. This phase transition is irreversible upon decompression. The second order bulk moduli of silicalite A and silicalite B, calcd. after the fulfillment of the phase transition, are: K 0=18.2(2) and K 0=14.3 (2) GPa, resp. These values makes silicalite the most compressible zeolite among those up to now studied in silicone oil. The structural deformations induced by HP in silicalite A were investigated by means of complete Rietveld structural refinements, before and after the phase transition, at P amb and 0.9 GPa, resp. The elastic behaviors of the three MFI-type zeolites here investigated were compared with those of Na-ZSM-5 and H-ZSM-5, studied in similar exptl. conditions: the two silicalites - which are the phases with the highest Si/Al ratios and hence the lowest extraframework contents - show the highest compressibility. On the contrary, the most rigid material is mutinaite, which has a very complex extraframework compn. characterized by a high no. of cations and water mols.
- 30Kokotailo, G. T.; Lawton, S. L.; Olson, D. H.; Meier, W. M. Structure of Synthetic Zeolite ZSM-5 Nature 1978, 272, 437– 438 DOI: 10.1038/272437a030Structure of synthetic zeolite ZSM-5Kokotailo, G. T.; Lawton, S. L.; Olson, D. H.; Meier, W. M.Nature (London, United Kingdom) (1978), 272 (5652), 437-8CODEN: NATUAS; ISSN:0028-0836.The structure of synthetic zeolite ZSM-5 was detd. The crystals are orthorhombic, space group Pnma, with a 20.1, b 19.9, and c 13.4 Å. The unit contents of the Na form are NanAlnSi96-nO192.∼16H2O with n <27 and typically ∼3. The ZSM-5 framework contains 2 intersecting channel systems, 1 sinusoidal parallel to [001] and 1 straight parallel to [010]. The elliptical 10-membered ring openings controlling the channels have an effective diam. between those of zeolite Linde A and faujasite.
- 31Nosé, S. A Unified Formulation of the Constant Temperature Molecular Dynamics Methods J. Chem. Phys. 1984, 81, 511 DOI: 10.1063/1.44733431A unified formulation of the constant-temperature molecular-dynamics methodsNose, ShuichiJournal of Chemical Physics (1984), 81 (1), 511-19CODEN: JCPSA6; ISSN:0021-9606.Three recently proposed const. temp. mol. dynamics methods [N., (1984) (1); W. G. Hoover et al., (1982) (2); D. J. Evans and G. P. Morris, (1983) (2); and J. M. Haile and S. Gupta, 1983) (3)] are examd. anal. via calcg. the equil. distribution functions and comparing them with that of the canonical ensemble. Except for effects due to momentum and angular momentum conservation, method (1) yields the rigorous canonical distribution in both momentum and coordinate space. Method (2) can be made rigorous in coordinate space, and can be derived from method (1) by imposing a specific constraint. Method (3) is not rigorous and gives a deviation of order N-1/2 from the canonical distribution (N the no. of particles). The results for the const. temp.-const. pressure ensemble are similar to the canonical ensemble case.
- 32Nosé, S. Constant Temperature Molecular Dynamics Methods Prog. Theor. Phys. Suppl. 1991, 103, 1– 46 DOI: 10.1143/PTPS.103.132Constant-temperature molecular-dynamics methodsNose, ShuichiProgress of Theoretical Physics Supplement (1991), 103 (), 1-46CODEN: PTPSEP; ISSN:0375-9687.A review with 62 refs. on how the canonical distribution is realized in simulations based on deterministic dynamical equations. Basic formulations and extensions of two const.-temp. mol. dynamics methods, the constrained and the extended system methods, are discussed.
- 33Bylander, D. M.; Kleinman, L. Energy Fluctuations Induced by the Nosé Thermostat Phys. Rev. B: Condens. Matter Mater. Phys. 1992, 46, 13756– 13761 DOI: 10.1103/PhysRevB.46.1375633Energy fluctuations induced by the Nose thermostatBylander; KleinmanPhysical review. B, Condensed matter (1992), 46 (21), 13756-13761 ISSN:0163-1829.There is no expanded citation for this reference.
- 34Momma, K.; Izumi, F. VESTA 3 for Three-Dimensional Visualization of Crystal, Volumetric and Morphology Data J. Appl. Crystallogr. 2011, 44, 1272– 1276 DOI: 10.1107/S002188981103897034VESTA 3 for three-dimensional visualization of crystal, volumetric and morphology dataMomma, Koichi; Izumi, FujioJournal of Applied Crystallography (2011), 44 (6), 1272-1276CODEN: JACGAR; ISSN:0021-8898. (International Union of Crystallography)VESTA is a 3D visualization system for crystallog. studies and electronic state calcns. It was upgraded to the latest version, VESTA 3, implementing new features including drawing the external morphpol. of crysals; superimposing multiple structural models, volumetric data and crystal faces; calcn. of electron and nuclear densities from structure parameters; calcn. of Patterson functions from the structure parameters or volumetric data; integration of electron and nuclear densities by Voronoi tessellation; visualization of isosurfaces with multiple levels, detn. of the best plane for selected atoms; an extended bond-search algorithm to enable more sophisticated searches in complex mols. and cage-like structures; undo and redo is graphical user interface operations; and significant performance improvements in rendering isosurfaces and calcg. slices.
- 35Tuckerman, M.; Laasonen, K.; Sprik, M.; Parrinello, M. Ab Initio Molecular Dynamics Simulation of the Solvation and Transport of Hydronium and Hydroxyl Ions in Water J. Chem. Phys. 1995, 103, 150 DOI: 10.1063/1.46965435Ab initio molecular dynamics simulation of the solvation and transport of hydronium and hydroxyl ions in waterTuckerman, M.; Laasonen, K.; Sprik, M.; Parrinello, M.Journal of Chemical Physics (1995), 103 (1), 150-61CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)Charge defects in water created by excess or missing protons appear in the form of solvated hydronium H3O+ and hydroxyl OH- ions. Using the method of ab initio mol. dynamics, we have investigated the structure and proton transfer dynamics of the solvation complexes, which embed the ions in the network of hydrogen bonds in the liq. In our ab initio mol. dynamics approach, the interat. forces are calcd. each time step from the instantaneous electronic structure using d. functional methods. All hydrogen atoms, including the excess proton, are treated as classical particles with the mass of a deuterium atom. For the H3O+ ion we find a dynamic solvation complex, which continuously fluctuates between a (H5O2)+ and a (H9O4)+ structure as a result of proton transfer. The OH- has a predominantly planar fourfold coordination forming a (H9O5)- complex. Occasionally this complex is transformed in a more open tetrahedral (H7H4)- structure. Proton transfer is obsd. only for the more waterlike (H7O4)- complex. Transport of the charge defects is a concerted dynamical process coupling proton transfer along hydrogen bonds and reorganization of the local environment. The simulation results strongly support the structural diffusion mechanism for charge transport. In this model, the entire structure-and not the constituent particles-of the charged complex migrates through the hydrogen bond network. For H3O+, we propose that transport of the excess proton is driven by coordination fluctuations in the first solvation shell (i.e., second solvation shell dynamics). The rate-limiting step for OH- diffusion is the formation of the (H7O4)- structure, which is the solvation state showing proton transfer activity.
- 36Henkelman, G.; Arnaldsson, A.; Jónsson, H. A Fast and Robust Algorithm for Bader Decomposition of Charge Density Comput. Mater. Sci. 2006, 36, 354– 360 DOI: 10.1016/j.commatsci.2005.04.010There is no corresponding record for this reference.
- 37Sanville, E.; Kenny, S. D.; Smith, R.; Henkelman, G. Improved Grid-Based Algorithm for Bader Charge Allocation J. Comput. Chem. 2007, 28, 899– 908 DOI: 10.1002/jcc.2057537Improved grid-based algorithm for Bader charge allocationSanville, Edward; Kenny, Steven D.; Smith, Roger; Henkelman, GraemeJournal of Computational Chemistry (2007), 28 (5), 899-908CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)An improvement to the grid-based algorithm of Henkelman et al. for the calcn. of Bader vols. is suggested, which more accurately calcs. at. properties as predicted by the theory of Atoms in Mols. The CPU time required by the improved algorithm to perform the Bader anal. scales linearly with the no. of interat. surfaces in the system. The new algorithm corrects systematic deviations from the true Bader surface, calcd. by the original method and also does not require explicit representation of the interat. surfaces, resulting in a more robust method of partitioning charge d. among atoms in the system. Applications of the method to some small systems are given and it is further demonstrated how the method can be used to define an energy per atom in ab initio calcns.
- 38Tang, W.; Sanville, E.; Henkelman, G. A Grid-Based Bader Analysis Algorithm without Lattice Bias J. Phys.: Condens. Matter 2009, 21, 084204 DOI: 10.1088/0953-8984/21/8/08420438A grid-based Bader analysis algorithm without lattice biasTang, W.; Sanville, E.; Henkelman, G.Journal of Physics: Condensed Matter (2009), 21 (8), 084204/1-084204/7CODEN: JCOMEL; ISSN:0953-8984. (Institute of Physics Publishing)A computational method for partitioning a charge d. grid into Bader vols. is presented which is efficient, robust, and scales linearly with the no. of grid points. The partitioning algorithm follows the steepest ascent paths along the charge d. gradient from grid point to grid point until a charge d. max. is reached. In this paper, we describe how accurate off-lattice ascent paths can be represented with respect to the grid points. This improvement maintains the efficient linear scaling of an earlier version of the algorithm, and eliminates a tendency for the Bader surfaces to be aligned along the grid directions. As the algorithm assigns grid points to charge d. maxima, subsequent paths are terminated when they reach previously assigned grid points. It is this grid-based approach which gives the algorithm its efficiency, and allows for the anal. of the large grids generated from plane-wave-based d. functional theory calcns.
- 39Wernet, P.; Nordlund, D.; Bergmann, U.; Cavalleri, M.; Odelius, M.; Ogasawara, H.; Näslund, L. A.; Hirsch, T. K.; Ojamäe, L.; Glatzel, P. The Structure of the First Coordination Shell in Liquid Water Science 2004, 304, 995– 999 DOI: 10.1126/science.109620539The Structure of the First Coordination Shell in Liquid WaterWernet, Ph.; Nordlund, D.; Bergmann, U.; Cavalleri, M.; Odelius, M.; Ogasawara, H.; Naeslund, L. A.; Hirsch, T. K.; Ojamaee, L.; Glatzel, P.; Pettersson, L. G. M.; Nilsson, A.Science (Washington, DC, United States) (2004), 304 (5673), 995-999CODEN: SCIEAS; ISSN:0036-8075. (American Association for the Advancement of Science)X-ray absorption spectroscopy and x-ray Raman scattering were used to probe the mol. arrangement in the 1st coordination shell of liq. H2O. The local structure is characterized by comparison with bulk and surface of ordinary hexagonal ice Ih and with calcd. spectra. Most mols. in liq. H2O are in 2 H-bonded configurations with 1 strong donor and 1 strong acceptor H bond in contrast to the 4 H-bonded tetrahedral structure in ice. Upon heating from 25° to 90°, 5 to 10% of the mols. change from tetrahedral environments to 2 H-bonded configurations. Findings are consistent with neutron and x-ray diffraction data, and combining the results sets a strong limit for possible local structure distributions in liq. H2O. Serious discrepancies with structures based on current mol. dynamics simulations are obsd.
- 40Ambrosetti, A.; Alfè, D.; Robert, A.; DiStasio, J.; Tkatchenko, A. Hard Numbers for Large Molecules: Toward Exact Energetics for Supramolecular Systems J. Phys. Chem. Lett. 2014, 5, 849– 855 DOI: 10.1021/jz402663k40Hard Numbers for Large Molecules: Toward Exact Energetics for Supramolecular SystemsAmbrosetti, Alberto; Alfe, Dario; DiStasio, Robert A.; Tkatchenko, AlexandreJournal of Physical Chemistry Letters (2014), 5 (5), 849-855CODEN: JPCLCD; ISSN:1948-7185. (American Chemical Society)Noncovalent interactions are ubiquitous in mol. and condensed-phase environments, and hence a reliable theor. description of these fundamental interactions could pave the way toward a more complete understanding of the microscopic underpinnings for a diverse set of systems in chem. and biol. In this work, we demonstrate that recent algorithmic advances coupled to the availability of large-scale computational resources make the stochastic quantum Monte Carlo approach to solving the Schrodinger equation an optimal contender for attaining "chem. accuracy" (1 kcal/mol) in the binding energies of supramol. complexes of chem. relevance. To illustrate this point, we considered a select set of seven host-guest complexes, representing the spectrum of noncovalent interactions, including dispersion or van der Waals forces, π-π stacking, hydrogen bonding, hydrophobic interactions, and electrostatic (ion-dipole) attraction. A detailed anal. of the interaction energies reveals that a complete theor. description necessitates treatment of terms well beyond the std. London and Axilrod-Teller contributions to the van der Waals dispersion energy.
- 41Huang, S.-J.; Yang, C.-Y.; Zheng, A.; Feng, N.; Yu, N.; Wu, P.-H.; Chang, Y.-C.; Lin, Y.-C.; Deng, F.; Liu, S.-B. New Insights into Keggin-Type 12-Tungstophosphoric Acid from 31P MAS NMR Analysis of Absorbed Trimethylphosphine Oxide and DFT Calculations Chem. - Asian J. 2011, 6, 137– 148 DOI: 10.1002/asia.20100057241New Insights into Keggin-Type 12-Tungstophosphoric Acid from 31P MAS NMR Analysis of Absorbed Trimethylphosphine Oxide and DFT CalculationsHuang, Shing-Jong; Yang, Chih-Yi; Zheng, Anmin; Feng, Ningdong; Yu, Ningya; Wu, Pei-Hao; Chang, Yu-Chi; Lin, Ying-Chih; Deng, Feng; Liu, Shang-BinChemistry - An Asian Journal (2011), 6 (1), 137-148CODEN: CAAJBI; ISSN:1861-4728. (Wiley-VCH Verlag GmbH & Co. KGaA)The acid and transport properties of the anhyd. Keggin-type 12-tungstophosphoric acid (H3PW12O40; HPW) have been studied by solid-state 31P magic-angle spinning NMR of absorbed trimethylphosphine oxide (TMPO) in conjunction with DFT calcns. Accordingly, 31P NMR resonances arising from various protonated complexes, such as TMPOH+ and (TMPO)2H+ adducts, could be unambiguously identified. It was found that thermal pretreatment of the sample at elevated temps. (≥423 K) is a prerequisite for ensuring complete penetration of the TMPO guest probe mol. into HPW particles. Transport of the TMPO absorbate into the matrix of the HPW adsorbent was found to invoke a desorption/absorption process assocd. with the (TMPO)2H+ adducts. Consequently, three types of protonic acid sites with distinct superacid strengths, which correspond to 31P chem. shifts of 92.1, 89.4, and 87.7 ppm, were obsd. for HPW samples loaded with less than three mols. of TMPO per Keggin unit. Together with detailed DFT calcns., these results support the scenario that the TMPOH+ complexes are assocd. with protons located at three different terminal oxygen (Od) sites of the PW12O403- polyanions. Upon increasing the TMPO loading to >3.0 mols. per Keggin unit, abrupt decreases in acid strength and the corresponding structural variations were attributed to the change in secondary structure of the pseudoliquid phase of HPW in the presence of excessive guest absorbate.