Unraveling the Role of H2 and NH3 in the Amination of Isohexides over a Ru/C CatalystClick to copy article linkArticle link copied!
- Hang HuHang HuEco-Efficient Products and Processes Laboratory (E2P2L), UMI 3464 CNRS-Solvay, 3966 Jin Du Road, Xin Zhuang Ind. Zone, 201108 Shanghai, ChinaIC2MP UMR CNRS_Université de Poitiers 7285, ENSIP 1 rue Marcel Doré, TSA 41195, 86073 Poitiers Cedex 9, FranceMore by Hang Hu
- Muhammad Akif RamzanMuhammad Akif RamzanUniversité de Lyon, Ecole Normale Supérieure de Lyon, CNRS UMR 5182, Laboratoire de Chimie, 46 allée d’Italie, F69364 Lyon, FranceMore by Muhammad Akif Ramzan
- Raphael Wischert*Raphael Wischert*Email: [email protected]Eco-Efficient Products and Processes Laboratory (E2P2L), UMI 3464 CNRS-Solvay, 3966 Jin Du Road, Xin Zhuang Ind. Zone, 201108 Shanghai, ChinaMore by Raphael Wischert
- François JerômeFrançois JerômeIC2MP UMR CNRS_Université de Poitiers 7285, ENSIP 1 rue Marcel Doré, TSA 41195, 86073 Poitiers Cedex 9, FranceMore by François Jerôme
- Carine Michel*Carine Michel*Email: [email protected]Université de Lyon, Ecole Normale Supérieure de Lyon, CNRS UMR 5182, Laboratoire de Chimie, 46 allée d’Italie, F69364 Lyon, FranceMore by Carine Michel
- Karine de Oliveira Vigier*Karine de Oliveira Vigier*Email: [email protected]IC2MP UMR CNRS_Université de Poitiers 7285, ENSIP 1 rue Marcel Doré, TSA 41195, 86073 Poitiers Cedex 9, FranceMore by Karine de Oliveira Vigier
- Marc Pera-Titus*Marc Pera-Titus*Email: [email protected]Eco-Efficient Products and Processes Laboratory (E2P2L), UMI 3464 CNRS-Solvay, 3966 Jin Du Road, Xin Zhuang Ind. Zone, 201108 Shanghai, ChinaCardiff Catalysis Institute, School of Chemistry, Cardiff University, Main Building, Park Place, CF10 3AT Cardiff, U.K.More by Marc Pera-Titus
Abstract
The direct amination of biomass-derived isohexides with NH3 over a Ru/C catalyst was systematically investigated to understand the role of H2 and NH3 in the production of isohexide diamines vs aminoalcohols, i.e., the transformation of one or both OH-groups in isohexides into NH2 groups. Only aminoalcohols with an exo-OH group were generated starting from isosorbide, which contains both an exo-OH and an endo-OH group, while a moderate yield of diamines was obtained from isomannide with two endo-OH groups due to the higher reactivity of the latter. The main byproducts were identified, including a variety of N- and O-containing cyclic compounds, such as 2,5-dimethylpyrrolidine, that arise from a decomposition path driven by hydrolysis/hydrodeoxygenation of a tricyclic amine intermediate. By combining density functional theory calculations with microkinetics, NH3 was found to adsorb strongly on the catalyst surface and generate adsorbed NH2 and NH species with variable coverage depending on the temperature and the nominal H2/NH3 ratio. Isomerization of isohexides was greatly suppressed by adsorbed NH3. Meanwhile, adsorbed NH3 discouraged the formation of byproducts driven by competing side reactions promoted by H2. The H2/NH3 ratio, which conditions the distribution of NH2 and NH species on the Ru surface, influences drastically the catalytic performance.
This publication is licensed under
License Summary*
You are free to share(copy and redistribute) this article in any medium or format and to adapt(remix, transform, and build upon) the material for any purpose, even commercially within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
*Disclaimer
This summary highlights only some of the key features and terms of the actual license. It is not a license and has no legal value. Carefully review the actual license before using these materials.
License Summary*
You are free to share(copy and redistribute) this article in any medium or format and to adapt(remix, transform, and build upon) the material for any purpose, even commercially within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
*Disclaimer
This summary highlights only some of the key features and terms of the actual license. It is not a license and has no legal value. Carefully review the actual license before using these materials.
License Summary*
You are free to share(copy and redistribute) this article in any medium or format and to adapt(remix, transform, and build upon) the material for any purpose, even commercially within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
*Disclaimer
This summary highlights only some of the key features and terms of the actual license. It is not a license and has no legal value. Carefully review the actual license before using these materials.
Synopsis
The amination of isohexides to diamines over Ru/C is affected by ammonia poisoning and formation of byproducts by hydrogenolysis.
Introduction
Results and Discussion
Influence of the Substrate on the Catalytic Performance
Analysis of the Byproducts Formed in the Amination of Isomannide
The most likely structures are circled by a dashed orange line. Unsaturation degree Ω = (C × 2 + 2 H + N)/2.
Role of NH3 and H2 in the Amination of Isomannide IM
Effect of NH3 and Aminoalcohols on the Amination of IM
Effect of H2 Pressure on the Amination of Isomannide IM
Reactivity of Ketones
Competition of Amination and Isomerization of Isomannide over Ru/C
Understanding the Role of NH3 and H2 in IM Amination
Conclusions
Experimental Section
Chemicals
General Procedure for Amination Reactions
General Procedure for the Synthesis of Isohexide endo-OH Monoketone
Identification of Products and Quantification
Computational Details
Supporting Information
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acssuschemeng.2c07501.
Coordinates for DFT calculations (ZIP)
Blank mass spectra of the MS system; mass spectra of AS2 and DAS; mass spectra of the reaction mixture after amination of isomannide IM under 30 bar H2; possible byproducts from IM amination; reductive amination of endo-OH ketone mixtures; optimized structures for different adsorption modes of aminoalcohols through either −OH or −NH2 functional groups; reaction profile corresponding to the dehydrogenation of NH3 to form NHx species on Ru(0001); GC methods A and B used for the analysis of product mixtures in the amination of IM; representative GC chromatograms obtained for reactant and product analysis using GC methods A and B; GC method C used for the analysis of product mixtures in the oxidation of IM and IS; representative GC chromatograms obtained for reactant and product analysis using GC methods A and C; steady-state coverage of surface species and empty catalytic sites as a function of temperature at a H2/NH3 ratio of 1.0; evolution of N2 gas as a function of temperature for H2/NH3 molar ratio of 1.0; and a table listing substrate conversion and product yield for figures in the main text (PDF)
Terms & Conditions
Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.
Acknowledgments
The authors are grateful to the Pôle Scientifique de Modélisation Numérique at the École Normale Supérieure de Lyon for HPC resources. The authors also thank the SYSPROD project and AXELERA Pôle de Compétitivité for financial support (PSMN Data Center). M.A.R. is grateful to the ENS de Lyon and Solvay for the financial support for his PhD grant. H.H. is also grateful to Solvay for funding a PhD grant.
DAM | diaminoisomannide |
DAS | diaminoisosorbide |
DAI | diaminoisoidide |
AM1 | aminoisomannide |
AM2 | exo-NH2-aminoisosorbide |
AS1 | exo-OH-aminoisosorbide |
AS1 | exo-OH-aminoisosorbide |
AS2 | aminoiidide |
II | isoidide |
IM | isomannide |
IS | isosorbide |
References
This article references 55 other publications.
- 1Wittcoff, H. A.; Reuben, B. G.; Plotkin, J. S. Industrial Organic Chemicals. 2nd ed., Wiley, NY, 2004.Google ScholarThere is no corresponding record for this reference.
- 2Lawrence, S. A. Amines: Synthesis, Properties and Applications; Cambridge University Press, 2004.Google ScholarThere is no corresponding record for this reference.
- 3Roose, P.; Eller, K.; Henkes, E.; Rossbacher, R.; Höke, H. Amines, Aliphatic, Ullmann’s Encyclopedia of Industrial Chemistry; Wiley-VCH: Weinheim, 2015, p 1.Google ScholarThere is no corresponding record for this reference.
- 4Imm, S.; Bahn, S.; Zhang, M.; Neubert, L.; Neumann, H.; Klasovsky, F.; Pfeffer, J.; Haas, T.; Beller, M. Improved ruthenium-catalyzed amination of alcohols with ammonia: synthesis of diamines and amino esters. Angew. Chem., Int. Ed. 2011, 50, 7599– 7603, DOI: 10.1002/anie.201103199Google Scholar4https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXosFemtL4%253D&md5=f5cdc49c0486d53561cb1b24a47660beImproved ruthenium-catalyzed amination of alcohols with ammonia: synthesis of diamines and amino estersImm, Sebastian; Baehn, Sebastian; Zhang, Min; Neubert, Lorenz; Neumann, Helfried; Klasovsky, Florian; Pfeffer, Jan; Haas, Thomas; Beller, MatthiasAngewandte Chemie, International Edition (2011), 50 (33), 7599-7603, S7599/1-S7599/11CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)A homogeneous selective catalytic amination of primary and secondary diols with ammonia is described. Catalyzed by [Ru(CO)ClH(PPh3)3]/Xantphos, the corresponding diamines were obtained in good yields using this method.
- 5Delidovich, I.; Hausoul, C.; Deng, L.; Pfützenreuter, R.; Rose, M.; Palkovits, R. Alternative Monomers Based on Lignocellulose and Their Use for Polymer Production. Chem. Rev. 2016, 116, 1540– 1599, DOI: 10.1021/acs.chemrev.5b00354Google Scholar5https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXhslGmsrbN&md5=202111f6f0d0f0517fc592e5e4c24f74Alternative Monomers Based on Lignocellulose and Their Use for Polymer ProductionDelidovich, Irina; Hausoul, Peter J. C.; Deng, Li; Pfuetzenreuter, Rebecca; Rose, Marcus; Palkovits, ReginaChemical Reviews (Washington, DC, United States) (2016), 116 (3), 1540-1599CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review. In recent years, the valorization of lignocellulose as feedstock for the prodn. of platform chems. has found tremendous attention. At the same time, the further transformation of these platform chems. into alternative monomers has been an area of growing research interest. Nevertheless, one has to point out that several of the monomers discussed in this review have already been synthesized by means of org. synthesis several decades ago. However, the renewed interest in lignocellulose based monomers entailed the design of green catalytic synthesis protocols to access these monomers.
- 6Pelckmans, M.; Renders, T.; Van de Vyver, S.; Sels, B. F. Bio-based amines through sustainable heterogeneous catalysis. Green Chem. 2017, 19, 5303– 5331, DOI: 10.1039/c7gc02299aGoogle Scholar6https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXhsFWiu77J&md5=cd61d96d3dd6031bfb81ead1ceb815c0Bio-based amines through sustainable heterogeneous catalysisPelckmans, M.; Renders, T.; Van de Vyver, S.; Sels, B. F.Green Chemistry (2017), 19 (22), 5303-5331CODEN: GRCHFJ; ISSN:1463-9262. (Royal Society of Chemistry)The prodn. of amines from biomass is a growing field of interest. Particularly the amination of bio-based alcs. receives a lot of attention. In this review, we discuss recent progress in the development of efficient heterogeneous catalysts. The substrate scope for the prodn. of bio-based amines is not limited to (hemi)cellulosic alcs. Other platform chems. that originate from different biomass fractions, such as lignin, oils, chitin and protein, are also suitable feedstock for the prodn. of amines. This comprehensive review first provides an overview of the available bio-based feedstock candidates. The following section is devoted to the sustainable reaction routes that are available to carry out the desired amination reactions. Next, state-of-the-art technologies are summarized for each substrate class, focussing on heterogeneous catalysis. Special attention is dedicated to the sustainability of the discussed reaction routes. Finally, a crit. discussion is provided, together with current challenges and future perspectives regarding the industrial prodn. of bio-based amine chems.
- 7Pelckmans, M.; Vermandel, W.; Van Waes, F.; Moonen, K.; Sels, B. F. Low-temperature reductive aminolysis of carbohydrates to diamines and aminoalcohols by heterogeneous catalysis. Angew. Chem. Int. Ed 2017, 56, 14540– 14544, DOI: 10.1002/anie.201708216Google Scholar7https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXhs1CntLrE&md5=8737fa48cad71d03a8906ede255e3f4aLow-Temperature Reductive Aminolysis of Carbohydrates to Diamines and Aminoalcohols by Heterogeneous CatalysisPelckmans, Michiel; Vermandel, Walter; Van Waes, Frederik; Moonen, Kristof; Sels, Bert F.Angewandte Chemie, International Edition (2017), 56 (46), 14540-14544CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)Short amines, such as ethanolamines and ethylenediamines, are important compds. in today's bulk and fine chems. industry. Unfortunately, current industrial manuf. of these chems. relies on fossil resources and requires rigorous safety measures when handling explosive or toxic intermediates. Inspired by the elegant working mechanism of aldolase enzymes, a novel heterogeneously catalyzed process-reductive aminolysis-was developed for the efficient prodn. of short amines from carbohydrates at low temp. High-value bio-based amines contg. a bio-derived C2 carbon backbone were synthesized in one step with yields up to 87 C%, in the absence of a solvent and at a temp. below 405 K. A wide variety of available primary and secondary alkyl- and alkanolamines can be reacted with the carbohydrate to form the corresponding C2-diamine. The presented reductive aminolysis is therefore a promising strategy for sustainable synthesis of short, acyclic, bio-based amines.
- 8Froidevaux, V.; Negrell, C.; Caillol, S.; Pascault, J.-P.; Boutevin, B. Biobased Amines: From Synthesis to Polymers; Present and Future. Chem. Rev. 2016, 116, 14181– 14224, DOI: 10.1021/acs.chemrev.6b00486Google Scholar8https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XhslOmu7fM&md5=970ff058a08b3941d6d727e2ffec07bfBiobased Amines: From Synthesis to Polymers; Present and FutureFroidevaux, Vincent; Negrell, Claire; Caillol, Sylvain; Pascault, Jean-Pierre; Boutevin, BernardChemical Reviews (Washington, DC, United States) (2016), 116 (22), 14181-14224CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)Amines are key-intermediates in chem. industry due to their nucleophilic characteristic which confers a high reactivity to them. Thus, they are key-monomers for the synthesis of polyamides, polyureas, polyepoxydes···which are all of growing interest in automotive, aerospace, building or health applications. Despite a growing interest for biobased monomers and polymers, and particularly polyamides, it should be noticed that very few natural amines are available. Actually, there is only chitosan and poly(lysine). In this review, we present both fundamental and applied research on the synthesis of biobased primary and secondary amines with current available biobased resources. Their use is described as building block for material chem. Hence, we first recall some background on the synthesis of amines, including the reactivity of amines. Second we focus on the synthesis of biobased amines from all sorts of biomass, from carbohydrate, from terpenes, or from oleochem. sources. Third, because they need optimization and technol. developments, we discuss some examples of their use for the creation of biobased polymers. We conclude on the future of the synthesis of biobased amines and their use in different applications.
- 9Brun, N.; Hesemann, P.; Esposito, D. Expanding the biomass derived chemical space. Chem. Sci. 2017, 8, 4724– 4738, DOI: 10.1039/c7sc00936dGoogle Scholar9https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXmsFemtLw%253D&md5=8d9a951ab73aa9488961bc3e312330fbExpanding the biomass derived chemical spaceBrun, Nicolas; Hesemann, Peter; Esposito, DavideChemical Science (2017), 8 (7), 4724-4738CODEN: CSHCCN; ISSN:2041-6520. (Royal Society of Chemistry)A review. Biorefinery aims at the conversion of biomass and renewable feedstocks into fuels and platform chems., in analogy to conventional oil refinery. In the past years, the scientific community has defined a no. of primary building blocks that can be obtained by direct biomass decompn. However, the large potential of this "renewable chem. Space" to contribute to the generation of value added bio-active compds. and materials still remains unexplored. In general, biomass derived building blocks feature a diverse range of chem. functionalities. In order to be integrated into value-added compds., they require addnl. functionalization and/or covalent modification thereby generating secondary building blocks. The latter can be thus regarded as functional components of bio-active mols. or materials and represent an expansion of the renewable chem. space. This perspective highlights the most recent developments and opportunities for the synthesis of secondary biomass derived building blocks and their application to the prepn. of value added products.
- 10Liang, G.; Wang, A.; Li, L.; Xu, G.; Yan, N.; Zhang, T. Production of primary amines by reductive amination of biomass-derived aldehydes/ketones. Angew. Chem., Int. Ed. 2017, 56, 3050– 3054, DOI: 10.1002/anie.201610964Google Scholar10https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXitVKntbo%253D&md5=26b501e0c8dedcc757e9655af485d6e2Production of Primary Amines by Reductive Amination of Biomass-Derived Aldehydes/KetonesLiang, Guanfeng; Wang, Aiqin; Li, Lin; Xu, Gang; Yan, Ning; Zhang, TaoAngewandte Chemie, International Edition (2017), 56 (11), 3050-3054CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)Transformation of biomass into valuable nitrogen-contg. compds. is highly desired, yet limited success has been achieved. Here we report an efficient catalyst system, partially reduced Ru/ZrO2, which could catalyze the reductive amination of a variety of biomass-derived aldehydes/ketones in aq. ammonia. With this approach, a spectrum of renewable primary amines was produced in good to excellent yields. Moreover, we have demonstrated a two-step approach for prodn. of ethanolamine, a large-market nitrogen-contg. chem., from lignocellulose in an overall yield of 10 %. Extensive characterizations showed that Ru/ZrO2-contg. multivalence Ru assocn. species worked as a bifunctional catalyst, with RuO2 as acidic promoter to facilitate the activation of carbonyl groups and Ru as active sites for the subsequent imine hydrogenation.
- 11Flèche, G.; Huchette, M. Preparation, properties and chemistry. Starch/Stärke 1986, 38, 26– 30, DOI: 10.1002/star.19860380107Google Scholar11https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL28XktVKksL8%253D&md5=8096fd2618fb0f588ac5c2f2fb79b799Isosorbide. Preparation, properties and chemistryFleche, G.; Huchette, M.Starch/Staerke (1986), 38 (1), 26-30CODEN: STARDD; ISSN:0038-9056.Isosorbide (I) is obtained by dehydration of sorbitol and therefore can be considered as a valuable product from biomass. The acid-catalyzed reaction gives rise to different anhydro-compds., but also to polymer-like products. Kinetics of sorbitol decrease, followed with the help of HPLC, shows, remarkedly, the different reactions taking place during the dehydration. Physicochem. properties of isosorbide are also discussed: melting temp., sp. gr., soly. Emphasis is put on stereochem. aspect, pointing out the endo-exo position of hydroxyl groups vs. those of other isomers: exo-exo for isoiodide and endo-endo for isomannide. Some other properties such as: hygroscopicity ant thermal stability are also discussed. The chem. reactivity of the mols. is described and some reactions analyzed, proving the interest of the cyclic conformation. Finally, known applications are presented.
- 12Wiggins, L. F. 2. The anhydrides of polyhydric alcohols. Part I. The constitution of isomannide. J. Chem. Soc. 1945, 4– 7, DOI: 10.1039/jr9450000004Google Scholar12https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaH2MXhslGlsg%253D%253D&md5=c6aa485bc95f2a1019a360b1527918daAnhydrides of polyhydric alcohols. I. Constitution of isomannideWiggins, L. F.Journal of the Chemical Society (1945), (), 4-7CODEN: JCSOA9; ISSN:0368-1769.Mannitol (I) (20 g.) in 70 cc. dichloroglycerol, ClCH2CH(OH)CH2Cl heated at 140-50° while HCl is passed through the suspension for 3.5 hrs., gives 5.4 g. of 1,4,3,6-dianhydromannitol (isomannide) (II), m. 87-8°, [α]D 91° (H2O, c 2.28). I (500 g.) in 3 l. fuming HCl, boiled 72 hrs., gives 140 g. of II. 1,6-Dichloromannitol, ClCH2[CH(OH)]4CH2Cl (III) (C.A. 38, 2628.6) (2 g.), heated in vacuo at 180°, gives 0.95 g. of II. II (4 g.) in 5.6 g. C5H5N, treated with SOCl2 at 0° and heated 0.5 hr. at 100°, gives 70% of the 2,5-di-Cl compd. (IV), m. 67°, [α]D 93.5° (CHCl3, c 2.054); this is unchanged on distn. with fused KOH. II (1 g.), treated 4 times with MeI and Ag2O at 45°, gives 0.9 g. of the 2,5-di-Me ether (IVA), m. 75-6°, [α]D18 175° (CHCl3, c 2.287). II (1 g.) in 20 cc. fuming HCl, heated 24 hrs. in a sealed tube, gives 0.25 g. of III; thus positions 1 and 6 are involved in the anhydro rings. II (1 g.), heated at 120° for 29 hrs. with 50 cc. MeOH satd. with NH3 at 0°, gives 0.8 g. unchanged II; it is also unaffected with 10% MeOH-MeONa under the same conditions. II is not oxidized by HNO3 (d. 1.14) on heating at 100° for 4 hrs. Pb(OAc)4 in AcOH is without action on II in 48 hrs., indicating that the 2 OH groups are not on adjacent C atoms. IV (7.2 g.) in 200 cc. fuming HCl, heated in a sealed tube for 72 hrs. at 110°, gives 1.6 g. of unchanged IV and 3.4 g. of 1,2,5,6-tetrachloromannitol (V), m. 69-70°, [α]D 28.3° (CHCl3, c 3.107); 3,4-di-Bz deriv., m. 109-10°, [α]D17 -95.4° (CHCl3, c 1.048). V and PCl5 at 130° for 1 hr. give hexachloromannitol; thus V possesses the configuration of I. With MeONa at room temp. for 4 hrs. 1 g. of V gives 0.5 g. of IV. V (0.3 g.), shaken overnight with 25 cc. Me2CO contg. 0.1 cc. H2SO4, gives 0.33 g. of the 3,4-acetone deriv. (VI), b0.05 115° (bath temp.), nD16 1.4954, [α]D 56.8° (CHCl3, c 2.405); VI results also in 0.4-g. yield from 1 g. of 1,6-dichloro-3,4-acetonemannitol (VII)(Micheel, C.A. 26, 4304) and SOCl2 in C5H5N (1 hr. at 100°). VII (5 g.), treated 5 times with MeI and Ag2O at 45° for 9 hrs., gives 1.6 g. of the 2,5-di-Me ether (VIII), m. 56°, [α]D 11.5° (CHCl3, c 2.253); 0.503 g. of VIII in 40 cc. of 75% EtOH contg. 5% H2SO4, kept at room temp. for 550 hrs. (change in [α]D from 9.5° to -35°), gives 0.21 g. of 1,6-dichloro-2,5-dimethylmannitol, m. 131°; with MeONa in MeOH 0.17 g. gives 0.09 g. of IVA; this shows that the HO groups of II are located at C2 and C5. Mannitan (IX) distils at 10 mm. practically without decompn.; distn. of 2 g. of IX with 2 drops of concd. H2SO4 at 10 mm. gives 1.1 g. of II. IX (5 g.) in 50 cc. C5H5N, treated at 0° with 6 g. p-MeC6H4SO2Cl and kept at room temp. for 96 hrs. (with addn. of 25 cc. Ac2O after 48 hrs.), gives 10.5 g. of 1-tosyl-2,4,5-triacetylmannitan (X), a liquid. Reaction of X with MeONa gives II. That the tosyl group is on a primary alc. group is shown by treatment of X with NaI in Me2CO to give 80% of p-MeC6H4SO3Na. The above facts prove that the rings are hydrofuranol in type.
- 13Rose, M.; Palkovits, R. Isosorbide as a renewable platform chemical for versatile applications – quo vadis?. ChemSusChem 2012, 5, 167– 176, DOI: 10.1002/cssc.201100580Google Scholar13https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XitVamsQ%253D%253D&md5=e5a646b04db3b90495d4256201d3be04Isosorbide as a Renewable Platform chemical for Versatile Applications-Quo Vadis?Rose, Marcus; Palkovits, ReginaChemSusChem (2012), 5 (1), 167-176CODEN: CHEMIZ; ISSN:1864-5631. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. Isosorbide is a platform chem. of considerable importance for the future replacement of fossil resource-based products. Applications as monomers and building blocks for new polymers and functional materials, are conceivable. This minireview deals with all aspects of isosorbide chem., which includes its prodn., special properties, and chem. transformations for its utilization in biogenic polymers and other applications of interest.
- 14Fenouillot, F.; Rousseau, A.; Colomines, G.; Saint-Loup, R.; Pascault, J. P. Polymers from renewable 1,4:3,6-dianhydrohexitols (isosorbide, isomannide and isoidide): a review. Progr. Polym. Sci. 2010, 35, 578– 622, DOI: 10.1016/j.progpolymsci.2009.10.001Google Scholar14https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXjvFKitr4%253D&md5=0ca3e0de11d9077e508554de0e4ac596Polymers from renewable 1,4:3,6-dianhydrohexitols (isosorbide, isomannide and isoidide): A reviewFenouillot, F.; Rousseau, A.; Colomines, G.; Saint-Loup, R.; Pascault, J.-P.Progress in Polymer Science (2010), 35 (5), 578-622CODEN: PRPSB8; ISSN:0079-6700. (Elsevier Ltd.)A review. The use of the 1,4:3,6-dianhydrohexitols isosorbide, isomannide, and isoidide in polymers is reviewed. 1,4:3,6-Dianhydrohexitols are derived from renewable resources from cereal-based polysaccharides. 1,4:3,6-Dianhydrohexitols are derived from renewable resources from cereal-based polysaccharides. In the field of polymeric materials, these diols are essentially employed to synthesize or modify polycondensates. Their attractive features as monomers are linked to their rigidity, chirality, non-toxicity, and the fact that they are not derived from petroleum. First, the synthesis of high glass transition temp. polymers with good thermomech. resistance is possible. Second, the chiral nature of 1,4:3,6-dianhydrohexitols may lead to specific optical properties. Finally, biodegradable polymers can be obtained. The prodn. of isosorbide on an industrial scale with a purity satisfying the requirements for polymer synthesis suggests that isosorbide will soon emerge in industrial polymer applications. However, a deciding factor will be the redn. of polymn. time of these low-reactivity monomers to values compatible with economically viable prodn. processes to give polyesters, polyamides, poly(amide esters), poly(ester imides), polycarbonates, polyurethanes, and polyethers.
- 15Wu, J.; Jasinska-Walc, L.; Dudenko, D.; Rozanski, A.; Hansen, M. R.; van Es, D.; Koning, C. E. An investigation of polyamides based on isoidide-2,5-dimethyleneamine as a green rigid building block with enhanced reactivity. Macromolecules 2012, 45, 9333– 9346, DOI: 10.1021/ma302126bGoogle Scholar15https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38Xhs12mtLrE&md5=7322665bbe1faba162bad35a629950adAn Investigation of Polyamides Based on Isoidide-2,5-dimethyleneamine as a Green Rigid Building Block with Enhanced ReactivityWu, Jing; Jasinska-Walc, Lidia; Dudenko, Dmytro; Rozanski, Artur; Hansen, Michael Ryan; van Es, Daan; Koning, Cor E.Macromolecules (Washington, DC, United States) (2012), 45 (23), 9333-9346CODEN: MAMOBX; ISSN:0024-9297. (American Chemical Society)Novel, semicryst. polyamides and copolyamides were synthesized from a new carbohydrate-based diamine, namely isoidide-2,5-dimethyleneamine (IIDMA). In combination with 1,6-hexamethylene diamine (1,6-HDA) as well as the biobased sebacic acid (SA) or brassylic acid (BrA), the desired copolyamides were obtained via melt polymn. of the nylon salts followed by a solid-state polycondensation (SSPC) process. Depending on the chem. compns., the no. av. mol. wts. (Mn) of the polyamides were in the range of 4000-49000 g/mol. With increasing IIDMA content in the synthesized copolyamides, their corresponding glass transition temps. (Tg) increased from 50 °C to approx. 60-67 °C while the melting temps. (Tm) decreased from 220 to 160 °C. The chem. structures of the polyamides were analyzed by NMR and FT-IR spectroscopy. Both differential scanning calorimetry (DSC) and wide-angle X-ray diffraction (WAXD) analyses revealed the semicryst. character of these novel copolyamides. Variable-temp. (VT) 13C{1H} cross-polarization/magic-angle spinning (CP/MAS) NMR and FT-IR techniques were employed to study the crystal structures as well as the distribution of IIDMA moieties over the cryst. and amorphous phases of the copolyamides. The performed ab initio calcns. reveal that the stability of the IIDMA moieties is due to a pronounced "boat" conformation of the bicyclic rings. The incorporation of methylene segments in between the isohexide group and the amide groups enables the hydrogen bonds formation and organization of the polymer chain fragments. Given the sufficiently high Tm values (∼200 °C) of the copolyamides contg. less than 50% of IIDMA, these biobased semicryst. copolyamides can be useful for engineering plastic applications.
- 16Montgomery, R.; Wiggins, L. F. 77. The anhydrides of polyhydric alcohols. Part IV. The constitution of dianhydro sorbitol. J. Chem. Soc. 1946, 390– 393, DOI: 10.1039/jr9460000390Google Scholar16https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaH28XisFKmug%253D%253D&md5=f1e77560ace47ba4790e450f6302f844Anhydrides of polyhydric alcohols. IV. Constitution of dianhydrosorbitolMontgomery, R.; Wiggins, L. F.Journal of the Chemical Society (1946), (), 390-3CODEN: JCSOA9; ISSN:0368-1769.cf. Hockett, et al., C.A. 40, 4677.3. Sorbitol (500 g.) and concd. HCl, refluxed 24 hrs., give 265 g. (66%) of 1,4,3,6-dianhydrosorbitol (I), b10 160-5°, m. 61-3°, [α]D 43.9° (H2O, c 0.8). I does not react with Pb(OAc)4 in AcOH or with BzH in the presence of ZnCl2. I (100 g.) in 200 cc. H2O at 40°, treated with 258 g. Me2SO4 and 360 cc. 30% NaOH in 10 portions during 2 hrs., gives 92 g. of the 2,5-di-Me deriv. (II), b0.1 93-5° (bath), nD19 1.4622, [α]D21 92.9° (CHCl3, c 1.51). I does not react with 5% MeONa in MeOH (20 hrs. at 120°) or with MeOH-NH3 (satd. at 0°) at 120° for 30 hrs. I (5 g.) and 50 cc. fuming HCl, heated at 100-10° for 24 hrs., give 5 g. of a dark brown sirup which, with (HCHO)3 and concd. H2SO4, yields 1,6-dichloro-2,4,3,5-dimethylenesorbitol, m. 116° (C.A. 38, 2628.6). 1,2-Acetoneglucose (40 g.) yields 30 g. of the 5,6-ditosyl deriv. (III) and 18.2 g. of the 3,5,6-tritosyl deriv., m. 129-30°, [α]D17 -3.4° (CHCl3, c 4.14), which results also on further tosylation of III (cf., however, Ohle, et al., C.A. 23, 103, who describe a 3,5,6-deriv. m. 95-6°, [α]D -5.2°). Reduction of 3,6-anhydro-glucose over Raney Ni at 110-20°/100 atm. gives 3,6-anhydrosorbitol (IV); distn. of 0.2 g. with a trace of H2SO4 gives 25 mg. of I. IV (1 g.) and 1.28 g. p-MeC6H4-SO2Cl in 10 cc. C5H5N, mixed at 0° and kept at room temp. for 60 hrs., Ac2O being added at 0° after the 1st 16 hrs., give 1.63 g. of 1-tosyl-2,4,5-triacetyl-3,6-anhydrosorbitol (V); 1.2 g. of V and 0.25 g. Na in 20 cc. MeOH and 10 cc. CHCl3, on standing overnight, give 73% of I. V (0.4 g.) and 0.29 g. NaI in 15 cc. Me2CO, heated at 110°, give 80% of p-MeC6H4SO3Na; the other product could not be crystd. Catalytic reduction of 1 g. of 2,5-dimethyl-3,6-anhydroglucose in H2O over 1 g. Raney Ni at 110-20°/100 atm. for 6 hrs., gives 1 g. of 2,5-dimethyl-3,6-anhydrosorbitol, m. 70-1°, [α]D22 -15.6° (CHCl3, c 0.173); treated as above with p-MeC6H4SO2Cl and Ac2O in C5H5N, 0.3 g. yields 0.23 g. of 1-tosyl-4-acetyl-2,5-dimethyl-3,6-anhydrosorbitol which, with MeONa in MeOH-CHCl3, gives 0.07 g. of II.
- 17Fletcher, H. G., Jr; Goepp, R. M., Jr Hexitol Anhydrides.1 1,4,3,6-Dianhydro-L-iditol and the Structures of Isomannide and Isosorbide. J. Am. Chem. Soc. 1946, 68, 939– 941, DOI: 10.1021/ja01210a007Google Scholar17https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaH28Xitl2jsQ%253D%253D&md5=b78bfdb5c19e6d0887e57a9ffd49914d1,4,3,6-Dianhydro-L-iditol and the structure of isomannide and isosorbideFletcher, Hewitt G., Jr.; Goepp, R. Max, Jr.Journal of the American Chemical Society (1946), 68 (), 939-41CODEN: JACSAT; ISSN:0002-7863.1,4,3,6-Dianhydrosorbitol (isosorbide) (I) (50 g.) and 10 g. Raney Ni in abs. EtOH, freed of EtOH in vacuo, heated 10 min. at 135-40°/2 mm., and then subjected to a pressure of 3750 lb./sq. in. of H at 190-200° for 2 hrs., gave a sirup which could not be crystd. Distn. of 46.1 g. at 140° (bath)/2 mm. gave 27.2 g. of distillate, [α]D25 44.2° in AcOH, and 18.1 g. of a residue (II), [α]D25 27.4° in AcOH. II (18.1 g.) in 50 ml. C5H5N at 0°, treated with 32 ml. BzCl, allowed to stand overnight at room temp., poured into 726 ml. ice water, and crystd. from EtOH, gave 31 g. of a product, m. 81.8-90°; the CHCl3, soln., shaken with aq. NaHCO3 to remove the BzOH and the product crystd. from EtOH and BuOH, gave 16.1 g. of the 2,5-di-Bz deriv. (III), m. 111-11.3° (m. ps. cor.), [α]D23.4 140.3° (CHCl3, c 2.03), [α]D28.2 110.5° (C5H5N, c 2.07), of 1,4,3,6-dianhydro-L-iditol (L-isoidide) (IV), m. 63.8-4.4°, [α]D24.5 20.8° (CHCl3, c 2.02), [α]D28.2 33.3° (C5H5N, c 2.24°). By the same method, 50 g. of isomannide (V) gives 0.92 g. of III. L-Iditol (0.97 g.) and 2 drops concd. H2SO4, heated at 140-5° (bath)/4 cm. for 1.25 hrs. and the product treated with BzCl in C5H5N, gave 31.4% of III. III (10 g.) in 40 ml. CHCl3 at 0°, treated with a chilled soln. of 0.1 g. Na in 40 ml. MeOH and kept at 0° for 23 hrs., gave 90.4% of IV, which is sparingly sol. in CHCl3. The prepn. of IV is best explained by the assumption that Raney Ni exerts a dehydrogenating action on the secondary alcs., converting 1 or both of the free HO groups to sym. CO groups, which are subsequently reduced with the formation of a mixt. of diastereoisomers. The isolation of IV from I and V proves the structure of IV and confirms those of I and V.
- 18Fletcher, H. G., Jr; Goepp, R. M., Jr 1,4;3,6-Hexitol dianhydride L-isoidide. J. Am. Chem. Soc. 1945, 67, 1042– 1043, DOI: 10.1021/ja01222a513Google Scholar18https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaH2MXitVKktw%253D%253D&md5=a4fed19fe54813233d8dcd51adc301491,4,3,6-Hexitol dianhydride. L-IsoidideFletcher, Hewitt G., Jr.; Goepp, R. Max, Jr.Journal of the American Chemical Society (1945), 67 (), 1042-3CODEN: JACSAT; ISSN:0002-7863.Reduction of D-isomannide or D-isosorbide at 200° over Raney Ni at 250 atms. gives a mixt. of hexitol dianhydrides from which, by benzoylation and fractional crystn., there is sepd. 1,4,3,6-dianhydro-L-iditol (L-isoidide) (I), m. 63.7-4.5°, [α]D24.5 20.8° (H2O, c 2.02), [α]D28.2 33.27° (C5H5N, c 2.24); dibenzoate, m. 111-11.3°, [α]D25.2 141.9° (CHCl3, c 2.15), [α]D28.2 110.5° (C5H5N, c 2.07). Similar data are given for the D-sorbitol and D-mannitol derivs. I also results by the direct acid-catalyzed anhydrization of L-iditol.
- 19Cope, C.; Shen, T. Y. The stereochemistry of 1,4: 3,6-dianhydrohexitol derivatives. J. Am. Chem. Soc. 1956, 78, 3177– 3182, DOI: 10.1021/ja01594a055Google Scholar19https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaG28Xnslyjsg%253D%253D&md5=1703546f35eedae5c3cca190ea364ca6Stereochemistry of 1,4:3,6-dianhydrohexitol derivativesCope, Arthur C.; Shen, T. Y.Journal of the American Chemical Society (1956), 78 (), 3177-82CODEN: JACSAT; ISSN:0002-7863.Ditosylate (I) (35 g.) of D-isomannide (II) and 40 g. Et4NOAc.H2O (III) refluxed 70 hrs. on the steam bath in 400 cc. Me2CO, concd. to 100 cc., dild. with 400 cc. H2O, and extd. with CHCl3, the extd. evapd., the residue distd., and the distillate (b0.7 120-50° bath) cooled gave 14.1 g. diacetate (IV) of 1,4:3,6-dianhydro-L-iditol (V), b0.5 100-10°, m. 57-7.6°, [α]D25 89.6° (c 1.5, CHCl3); all m.ps. are cor. II (4.0 g.) treated 2 days at 0-5° with 12 cc. Ac2O and 15 cc. pyridine, dild. with H2O, and extd. with CHCl3 gave 5.63 g. diacetate (VI) of 1,4:3,6-dianhydro-D-mannitol (VII), b0.5 118°, nD25 1.4680, [α]D25 194.5° (c 6.7, CHCl3). IV (12.5 g.) in 90 cc. abs. MeOH treated at 0° with 10 cc. N Ba(OMe)2 in MeOH, kept 18 hrs. at 5°, treated with powd. Dry Ice, and centrifuged, and the supernatant liquid worked up yielded 7.1 g. V, m. 43-3.5°, [α]D25 18.7° (c 2, H2O). IV (350 mg.) in 5 cc. 20% KOH shaken 20 min. with 1 cc. BzCl yielded 386 mg. dibenzoate of V, flakes, m. 110.6-11.4° (from aq. EtOH), [α]D26 134.2° (c 3, CHCl3). V (1.46 g.) treated 18 hrs. at 0-5° with 4.5 g. p-MeC6H4SO2Cl in 15 cc. pyridine and dild. with ice water yielded 3.96 g. di-p-toluene-sulfonate (VIII) of V, m. 105.5-106° (from MeOH), [α]D25 33.2° (c 2.5, CHCl3). Ditosylate (IX) (9.0 g.) of isosorbide refluxed 72 hrs. with 5.0 g. III in Me2CO, concd., dild. with H2O, and extd. with CHCl3 yielded 4.54 g. 5-O-acetyl-2-O-p-toluenesulfonyl-1,4:3,6-dianhydro-L-iditol (X), m. 95.5-6.3° (from MeOH), [α]D25 50.5° (c 4.7, CHCl3). X (1.0 g.) in 10 cc. abs. MeOH treated at 0° with 10 cc. 0.1N Ba(OMe)2 in MeOH, neutralized after 18 hrs. with powd. Dry Ice, centrifuged, and evapd. in vacuo, the oily residue treated 18 hrs. with 1.0 g. p-MeC6H4SO2Cl in 10 cc. pyridine at 5°, and the mixt. poured into ice water and filtered yielded 0.67 g. VIII, m. 105-6°. VIII (3.8 g.) and 1.75 g. III in 50 cc. Me2CO refluxed 72 hrs., concd. to 20 cc. and dild. with H2O gave 2.19 g. unchanged VIII, m. 104.5-5.3°. I (22.6 g.), 20 g. K phthalimide, and 300 cc. HCONMe2 heated 40 hrs. at 110° and dild. with 1.5 l. H2O, and the pptd. solid extd. with 300 cc. boiling EtOH left 6.9 g. 2,5-diphthalimido-2,5-dideoxy deriv. (XI) of V, prisms, m. 243.4-3.6° (from EtOAc-EtOH), [α]D25 168° (c 1, CHCl3). XI (8.08 g.), 1.51 g. 85% N2H4.H2O, and 200 cc. EtOH refluxed 2 days, treated with 20 cc. 4N HCl, refluxed 1 hr., and filtered, the filtrate evapd. to dryness in vacuo, the residue dissolved in 30 cc. H2O, filtered, treated with Darco, acidified to pH about 4.5 with aq. (CO2H)2, and dild. with EtOH to ppt. 2.1 g. oxalate, m. 242-3°, and the oxalate dissolved in H2O, treated with 0.7 g. NaOH in H2O, and distd. yielded 0.97 g. 2,5-di-NH2 analog (XII) of XI, b0.2 110°, m. 59-60°; picrate, m. 227.8-28.1°. I (39 g.) and 70 g. Me2NH shaken 72 hrs. at 120° in 500 cc. tetrahydrofuran in a steel bomb, the mixt. concd., treated with 100 cc. 20% aq. NaOH and extd. with Et2O, the ext. evapd., and the sirupy residue distd. yielded 7.5 g. 2,5-di-(Me2NCH2) analog (XIII) of XI, b0.2 80-100°, m. 57.5-8.5° (sublimed at 40° and 0.2 mm.), [α]D26 30.0° (c 2.1, H2O). XIII treated with excess MeI in boiling MeOH and recrystd. from aq. EtOH gave XIII.2MeI.H2O, m. 260° (decompn.), [α]D26 33.3° (c 2.2, H2O). XIII gave a dipicrate, m. 222° (decompn.) (from aq. EtOH). IX (45.2 g.) and 100 g. Me2NH in 600 cc. tetrahydrofuran heated at 120-30° with shaking in a bomb, cooled, concd., treated with 150 cc. 20% NaOH, and extd. with Et2O, and the ext. worked up gave 0.68 g. crude low boiling fraction, b0.3 41°, and 14.5 g. residue; the distillate treated with MeI in MeOH yielded 1.13 g. 5-dimethylamino-5-deoxy-1,4:3,6-dianhydro-1,2-L-iditoleen methiodide, plates, m. 202-3° (decompn.), [α]D31 33.1° (c 2.5, H2O); the nonvolatile residue in Et2O washed with H2O, dried over KOH, and concd. gave a brown sirup, which treated with picric acid yielded the picrate of 2-O-p-toluenesulfonyl-5-dimethylamino-5-deoxy-1,4:3,6-dianhydro-L-iditol (XIV), m. 179.4-80.8°. The sirup treated with MeI in MeOH gave XIV.MeI, prisms, m. 177.6-8.6° (from EtOH). IX (45.2 g.) and 100 g. Me2NH in 600 cc. tetrahydrofuran shaken 48 hrs. at 165°, concd., treated with 150 cc. 20% aq. NaOH, and extd. with Et2O, and the ext. worked up yielded 13.4 g. 2,5-bis(dimethylamino)-2,5-dideoxy-1,4:3,6-dianhydro-D-glucitol (XV), white prisms, m. 54.6-5.4° (sublimed at 80° and 0.5 mm.), [α]D25 106.3° (c 1.6, H2O). Cryst. oxalate (1.17 g.) of 2,5-di-NH2 analog (XVI) of XV in 10 cc. 98% HCO2H and 8 cc. 37% CH2O refluxed 18 hrs., treated with 10 cc. 4N HCl and evapd. to dryness in vacuo, the residue dissolved in 30 cc. N NaOH and extd. with Et2O, the ext. worked up, and the residue sublimed at 80° and 0.5 mm. gave 0.67 g. XV, m. 54-4.5°, which gave XV.2MeI.H2O, m. 292-5° (decompn.). XVI (3.0 g.) and 25 g. MeI in 100 cc. abs. MeOH refluxed 48 hrs. with stirring with 4 g. Na2CO3, acidified with 48% HI, and evapd. to dryness in vacuo yielded 9.25 g. XVI.2MeI, m. 289-94° (decompn.). XVI.2MeI in 100 cc. dry 1-methylmorpholine treated slowly with 2 g. LiAlH4 in 20 cc. 1-methylmorpholine with stirring and cooling, the mixt. heated 48 hrs. at 90-5°, treated with cooling with EtOAc, and centrifuged, the ppt. dissolved in 10% NaOH and extd. with Et2O, and the combined ext. and 1-methylmorpholine soln. fractionated gave 2.73 g. XV, m. 51.5-3.5°. XV (1.0 g.) warmed 15 min. on the steam bath with 2.0 g. p-MeC6H4SO2Cl in 20 cc. 10% aq. NaOH gave only 135 mg. pure XV, m. 54.3-5.3°. XII (0.72 g.) in 10 cc. 98% HCO2H and 6 cc. 37% CH2O refluxed 18 hrs. gave 0.72 g. XIII, m. 52-4°. XII (0.320 g.) and 5 g. MeI in 20 cc. abs. MeOH refluxed 65 hrs. with 0.85 g. NaHCO3 and worked up in the usual manner yielded 0.667 g. XIII.2MeI, prisms, m. 225-8° (decompn.) (from aq. EtOH). XII (0.72 g.) and 3 cc. concd. HCl in 30 cc. glacial AcOH treated slowly with cooling and stirring with 2.0 g. BuONO, the soln. kept 0.5 hr. at 0° and 18 hrs. at room temp., concd. in vacuo, dild. with H2O, and extd. with CHCl3, the ext. worked up, the residual sirup distd., the resulting oil chromatographed on 10 g. Al2O3 with 5:1 petr. ether-Et2O, and the 1st 60 cc. effluent worked up gave 103 g. XVII (R = R' = Cl), m. 67° XVI (0.72 g.) deaminated in the same manner yielded 87 mg. XVII. On the basis of these results D-isomannide has 2 endo-OH groups, isosorbide has 1 endo at C-5 and 1 exo at C-2, and L-isoidide has 2 exo-OH groups. Revised configurations are suggested for a no. of dianhydrohexitol derivs. and other compds. contg. bicyclic ring systems with 2 cisfused 5-membered rings on the basis of the steric effects observed (previous configuration and revised configuration given): 5-chloro-5-deoxy(?)-1,4:3,6-dianhydro-D-glucitol (W. G. Overend, et al., C.A. 43, 2939f), 5-chloro-5-deoxy-1,4:3,6-dianhydro-L-iditol; x-iodo-x'-O-p-toluenesulfonyl-1,4:3,6-dianhydro-D-glucitol (Hockett, et al., C.A. 40, 4677.3), 5-iodo-5-deoxy-2-O-p-toluenesulfonyl-1,4:3,6-dianhydro-L-iditol; 2,5-dichloro-2,5-dideoxy-1,4:3,6-dianhydro-D-mannitol (Wiggins, C.A. 39, 2736.9), 2,5-dichloro-2,5-dideoxy-1,4:3,6-dianhydro-L-iditol; 2-chloro-2-deoxy-5-O-methanesulfonyl-1,4:3,6-dianhydro-D-mannitol (Montgomery and Wiggins, C.A. 43, 2940a), 2-chloro-2-deoxy-5-O-methanesulfonyl-1,4:3,6-dianhydro-D-glucitol; 2-chloro-2-deoxy-5-phenylcarbamyl(?)-1,4:3,6-dianhydro-D-mannitol (Carr´e and Mauclere, C.A. 25, 4526), 2-chloro-2-deoxy-5-phenylcarbamyl-1,4:3,6-dianhydro-D-glucitol; 2,5-diiodo-2,5-dideoxy-1,4:3,6-dianhydro-D-mannitol (Hockett, et al., C.A. 40, 4677.8), 2,5-diiodo-2,5-dideoxy-1,4:3,6-dianhydro-L-iditol; 2,5-diamino-2,5-dideoxy-1,4:3,6-dianhydro-D-mannitol and derivs. (Montgomery and Wiggins, C.A. 40, 5016.4), 2,5-diamino-2,5-dideoxy-1,4:3,6-dianhydro-L-iditol and derivs.; 2,5-dithio-2,5-dideoxy-1,4:3,6-dianhydro-D-mannitol and derivs. (Bladon and Owen, C.A. 44, 6811h), 2,5-dithio-2,5-dideoxy-1,4:3,6-dianhydro-L-iditol and derivs.
- 20Brimacombe, J.; Foster, A. B.; Stacey, M.; Whiffen, D. H. Aspects of stereochemistry – I: Properties and reactions of some diols. Tetrahedron 1958, 4, 351– 360, DOI: 10.1016/0040-4020(58)80056-3Google Scholar20https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaG1MXmslGgsg%253D%253D&md5=002e6c00694f528888914553ddabe043Aspects of stereochemistry. I. Properties and reactions of diolsBrimacombe, J. S.; Foster, A. B.; Stacey, M.; Whiffen, D. H.Tetrahedron (1958), 4 (), 351-60CODEN: TETRAB; ISSN:0040-4020.The extent of intramol. H-bonding, as detd. by infrared spectroscopy in the HO stretching region in certain vicinal diols provided evidence for the stabilities of certain conformations. With the exception of cyclohexane-trans-1,2-diol (I), the carbocyclic trans diols were prepd. by bromination of the corresponding olefins, solvolysis of the trans bromides with AcOH and AgOAc and sapon. of the trans diacetates (Winstein and Roberts, C.A. 48, 5812f). The carbocyclic cis diols were obtained by hydroxylation of the appropriate olefins with KMnO4 (Clarke and Owen, C.A. 43, 7434h). Cycloheptene (12 g.) in 40 ml. Et2O at 0° brominated in Et2O, the colored soln. washed successively with 0.01N NaOH and water, the dried (MgSO4) soln. evapd. and the residue distd. gave 25.4 g. trans-1,2-dibromocycloheptane b20-5 138-40°, n23.5D 1.5532; this added to 40 g. AgOAc in 80 ml. AcOH and 25 ml. Ac2O (previously kept 2 hrs. at 110°), the mixt. kept 15 hrs. at 110° and the filtered soln. evapd. gave 8.2 g. trans-1,2-diacetoxycycloheptane b0.5 99-101°, n23D 1.4530. The diacetate (4 g.) boiled 2 hrs. in 8 ml. alc. and 8 ml. 35% aq. NaOH, the mixt. dild. with 8 ml. water and extd. 1 day with Et2O yielded 74% cycloheptane-trans-1,2-diol (II), b0.5 100°, m. 60-2°. Similarly were prepd. indan-trans-1,2-diol (III), m. 156-7°, and cyclopentane-trans-1,2-diol (IV), b15-20 131°. I, m. 102-3°, was prepd. according to Roebuck and Adkins [Org. Syntheses, Collective Vol. III, 217(1955)]. The cis-1,2-diols of cyclohexane (V), m. 93-4°, cycloheptane (VI), m. 45-6°, cyclopentane (VII), b29 133°, and indan (VIII), m. 96-8°, were obtained in 14, 11, 11, and 10.8-14.5% yields, resp. 3,4-Di-O-acetyl-D-xylal [3 g., m. 36-8°, [α]20D -288° (c 1.0, H2O)] prepd. according to Overend, et al. (C.A. 44, 6819f) in 100 ml. 1:1 alc.-H2O hydrogenated 30 min. with 100 mg. PtO2, the filtered soln. evapd. and the product distd. gave 1.3 g. 3,4-di-O-acetyl-1,5-anhydro-2-deoxy-D-threopentitol (IX), b0.5 102°, [α]20D -38° (c 0.85, CHCl2), [M]D -77°. IX (1.4 g.) refluxed 2 hrs. in 8 ml. 6N NaOH and 8 ml. alc., the soln. filtered through Amberlite IR-120 (H+) and the filtrate evapd. yielded 60.7% 1,5-anhydro-2-deoxy-D-threopentitol (X), b0.5 75-80°, m. 69°, [α]20D -29.6° (c 2.5, H2O), [M]D -35°. Hydrogenation of 3.2 g. 3,4-di-O-acetyl-L-arabinal (b0.01 110-40°, [α]20D -236° (c 1.12, CHCl3), prepd. according to Deriaz, et al. (C.A. 44, 2453a), and worked up similarly to X gave 2.2 g. 3,4-di-O-acetyl-1,5-anhydro-2-deoxy-L-erythropentitol (XI), b0.2 86-90°, [α]20D 75° (c 1.0, H2O), [M]D 151°. XI (1.4 g.) sapond. as for IX yielded 60.7% 1,5-anhydro-2-deoxy-L-erythropentitol (XII), b0.2-0.3 120°, [α]20D 64° (c 1.5, H2O), [M]D 75°. L-Arabinal (0.8 g., m. 80-2°, [α]20D -202° (c 3.0, H2O) prepd. according to D., et al.) in 20 ml. 1:1 alc.-H2O hydrogenated 1 hr. with 50 mg. PtO2 and the filtered soln. evapd. yielded 70.4% XII. DiBu L-tartrate (10 g., n16.5D 1.4450, [α]20D 10.2°) in 50 ml. tetrahydrofuran added slowly with stirring to 6 g. LiAlH4 in 150 ml. tetrahydrofuran and 75 ml. Et2O, the mixt. refluxed 1.5 hrs. and decompd. with 200 ml. H2O, the mixt. centrifuged and the solid residue washed with water, the combined centrifugates evapd. and the residue taken up in 200 ml. 1:1 H2O-MeOH, the soln. neutralized with CO2 and the filtered soln. evapd., the glassy solid taken up in 20 ml. 1:1 H2SO4-H2O and shaken 1 hr. with 15 ml. BzH, the mixt. dild. with water and filtered, the ppt. washed (water) and crystd. (PhMe) yielded 2.2 g. 1,2,3,4-di-O-benzylidene-L-threitol (Klosterman and Smith, C.A. 48, 2585f), m. 216-17°, [α]25D 81° (c 0.5, CHCl3), hydrolyzed 30 min. in 40 ml. 1:3 N H2SO4-alc., the hydrolyzate concd. and extd. with Et2O, the ext. filtered through Amberlite IRA-400 (HO-), the filtrate evapd. and the residue recrystd. to give 0.5 g. L-threitol (XIII), m. 88-9°, [α]20D -4° (c 8.0, H2O). XIII (2.17 g.) in 2.2 g. H2O and 2.2 g. H2SO4 heated 24 hrs. at 120° in a sealed tube, the hydrolyzate dild. with water, the soln. filtered through Amberlite IRA-400 (HO-) and the filtrate concd. gave 1.1 g. 1,4-anhydro-L-threitol (XIV), b15-17 120°, m. 63-4°, [α]20D -4° (c 7.2, H2O). Erythritol (3 g.) in 3 g. H2O and 3 g. H2SO4 heated 2 days at 120° and in a sealed tube and the hydrolyzate worked up as for XIII gave 1,4-anhydroerythritol (XV), b2-3 144°, n20D 1.4767. The zone electrophoretic (ionophoretic) mobility of the diols were detd. using the app. and technique of Foster, et al. (C.A. 50, 4584g), with a borate buffer (pH 10). The mobility, MG is defined as that mobility relative to D-glucose under standard conditions. The diols (0.43 millimole) in 5 ml. H2O were treated with 15 ml. 0.05M NaIO4 and the vol. rapidly adjusted to 100 ml. at 0°, aliquots withdrawn periodically titrated for unchanged NaIO4 by the standard arsenite method and where possible the times of half oxidation (t0.5) were recorded. The diols (26 mg.) in AcOH at 20° were treated with 49 ml. 0.1315N Pb(OAc)4 in AcOH at 20°, the vol. adjusted to 50 ml. and the unconsumed oxidant detd. on 5-ml. aliquots according to Hockett and McClenahan (C.A. 33, 68034). Methylation of 1,3-O-methylidene glyceritol (XVI), b11 82°, n20D 1.4533, according to Hibbert and Carter (C.A. 23, 98) gave 5-O-methyl-1,3-dioxan-5-ol, b. 147°, n25D 1.4230, converted by acid hydrolysis to 2-O-methyl-1,3-O-methylidene glyceritol, b13 120°, n23D 1.4476, completely resistant to attack by Pb(OAc)4 under the above conditions. Infrared spectra were measured in 2-cm. layers in CCl4 according to Spedding and Whiffen (C.A. 51, 15275b) with concns. of less than 0.005M diol. The results are tabulated (compd., ν for free and bonded OH in cm.-1, arithmetical difference between frequencies, arithmetical difference between standard secondary (3629 cm.-1) and bonded OH group frequencies, t0.5 for NaIO4 oxidation, t0.5 for Pb(OAc)4 oxidation and MG given): I, 3633, 3602, 31, 27, rapid, 1.9 hrs., 0.00; X, 3633, 3608, 25, 21, 20 min., 5 hrs., 0.00; V, 3632, 3592, 40, 37, rapid, 5 min., 0.07; XII, 3633, 3583, 50, 46, rapid, 9 min., 0.23; IV, 3624, -, -, -, 5 min., 10 min., 0.00; XIV, 3624, -, -, -, -, 11 hrs., 0.00; VII, 3624, 3579, 45, 50, rapid, rapid, 0.69; XV, 3624, 3585, 39, 44, -, rapid, 0.88; 1,4,3,6-dianhydro-D-glucitol, 3624, 3540, 84, 89, -, -, -; 1,4,3,6-dianhydro-D-glucitol, 3624, 3540, 84, 89, -, -, -; 1,4,3,6-dianhydro-D-glucitol, 3624, 3540, 84, 89, -, -, -; 1,4,3,6-dianhydro-D-mannitol, -, 3540, -, 89, -, -, -; 1,4,3,6-dianhydro-L-iditol, 3624, -, -, -, -, -, -; III, 3624, -, -, -, 60 min., -, 0.00; VII, 3620, 3579, 41, 50, rapid, -, 0.72; II, 3626, 3589, 37, 40, rapid, -, 0.53; VI, 3632, 3588, 44, 41, rapid, -, 0.69; XVI, 3625, 3593, 42, 36, -, -, -. In certain compds. the stabilities of different conformations can be affected by H-bonding from a substituent HO group to a ring O. Using the parameters evaluated by Whiffen (C.A. 51, 4261i) [M]D in aq. soln. were calcd. for the possible chair conformations of certain pyran derivs. (diol, conformation, [M]D calcd., and [M]D observed given): X, HO groups axial, -43°, -35°, HO groups equatorial, -45°, -; XII, C-3 HO group axial, -45°, -, C-4 HO group axial, 88°, 75°. The rate of reaction of the vicinal diols of these cyclic systems with glycol splitting reagents, and their zone electrophoretic mobility in an alk. borate buffer is influenced by the presence of a ring O atom.
- 21Montgomery, R.; Wiggins, L. F. 78. The anhydrides of polyhydric alcohols. Part V. 2 : 5-Diamino 1 : 4–3 : 6-dianhydro mannitol and sorbitol and their sulphanilamide derivatives. J. Chem. Soc. 1946, 0, 393– 396, DOI: 10.1039/jr9460000393Google Scholar21https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaH28XisFKmuw%253D%253D&md5=ff5e9a4f4c99a95b6626adda67069ca2Anhydrides of polyhydric alcohols. V. 2,5-Diamino-1,4,3,6-dianhydromannitol and -sorbitol and their sulfanilamide derivativesMontgomery, R.; Wiggins, L. F.Journal of the Chemical Society (1946), (), 393-6CODEN: JCSOA9; ISSN:0368-1769.1,4,3,6-Dianhydromannitol (59 g.) in 300 cc. dry C5H5N, treated at 0° with 155 g. p-MeC6H4SO2Cl in portions and allowed to warm to and stand at room temp. for 24 h., gives 162 g. of the 2,5-ditosyl deriv. (I), m. 93-4°, [α]D20 92.2° (CHCl3, c 2.582). I (50 g.) in 1500 cc. MeOH-NH3 (satd. at 0°), heated at 170-80° for 30 h., the residue heated with 42 g. Ba(OH)2 in 400 cc. H2O for 1 h., the residue from this operation freed from H2O by repeated distn. with C6H6, and extd. 6 times with 250-cc. portions of CHCl3, gives 12.8 g. of 2,5-diamino-1,4,3,6-dianhydromannitol (II), a sirup, b0.01 150° (bath), m. 59-62°, [α]D20 33.6° (CHCl3, c 2.322), which could be kept in cryst. condition only under N in sealed tubes; oxalate, m. 246-7° (decompn.); adipate, m. 189°; picrate, m. 227-8° (decompn.); sulfate, decomps. above 310°; dimethylenemucate, m. 246-7° (decompn.). II did not inhibit the growth of Staphylococcus aureus in vitro. II yields a disalicylidene deriv., yellow, m. 188-9°. The bis(p-acetamidophenylsulfonyl) deriv. (III), m. 278-9°, was prepd. from II and p-AcNHC6H4SO2Cl in dil. Me2CO contg. NaOH or NaHCO3 on stirring at room temp. for 45 min. or in C5H5N at room temp. for 2 days; it is optically inactive. Hydrolysis of 10 g. of III in 100 cc. Me2CO and 200 cc. 2 N HCl (refluxing 6 h.), or by heating 1 g. with 10 cc. 10% NaOH at 100° for 2 h., gives 2,5-disulfanilamido-1,4,3,6-dianhydromannitol (IV), m. 227-8°; it is optically inactive and does not form salts in aq. soln.; dry HCl in Me2CO-C6H6 (1:1) yields a di-HCl salt which effervesces between 180° and 215°, and yields IV with cold H2O. Reacetylation of IV (AcOH-Ac2O) gives III. II (3 g.) and p-O2NC6H4SO2Cl in C5H5N yield 9.3 g. of the bis(p-nitrophenylsulfonyl) deriv. (V), m. 213-14°, [α]D20 7.5° (Me2CO, c 2.65); redn. with Sn and HCl, or over Raney Ni, gives IV. Dianhydrosorbitol (58 g.) yields 190 g. of the 2,5-ditosyl deriv. (V), m. 101-2°, [α]D23 57.8° (CHCl3, c 4.945); treated as in the case of I, 50 g. of V yields 8.5 g. (54%) of 2,5-diamino-1,4,3,6-dianhydrosorbitol (VI), b0.01 105-10° (bath), nD18 1.5165, [α]D 43.6° (H2O, c 1.538); oxalate, m. 253-4° (decompn.); picrate, m. 200° (decompn.); HCl salt, does not m. 320°; sulfate, does not m. 330°; dimethylenemucate, m. 235-6° (decompn.); dimethylenesaccharate, m. 220-1° (decompn.). The salts did not inhibit the growth of S. aureus in vitro. The disalicylidene deriv. of VI m. 186-7°; bis(p-acetamidophenylsulfonyl) deriv., m. 263-4°, [α]D 51.4° (Me2CO-H2O (1:1), c 0.6); 2,5-disulfanilamido-1,4,3,6-dianhydrosorbitol (VII), m. 239-40°, [α]D 49.2° (Me2CO, c 0.406); the di-HCl salt, prepd. under anhyd. conditions, effervesces about 130° and is completely hydrolyzed by cold H2O. Bis(p-nitrophenylsulfonyl) deriv., m. 216-17°, [α]D20 56.8° (Me2CO, c 2.138); catalytic redn. over Raney Ni at room temp. and atm. pressure yields VII. IV and VII were only slightly sol. in H2O (0.02 g./100 cc.) but were fairly sol. in dil. HCl; they are inferior in bacteriostatic activity to sulfathiazole.
- 22Bashford, V. G.; Wiggins, L. F. 82. Anhydrides of polyhydric alcohols. Part XIII. The amino-derivatives of 1 : 4–3 : 6-dianhydro-mannitol, -sorbitol, and -L-iditol, and their behaviour towards nitrous acid. J. Chem. Soc. 1950, 0, 371– 374, DOI: 10.1039/jr9500000371Google Scholar22https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaG3cXisleiug%253D%253D&md5=3fd894628fd5232d88f72d0dbf4edb90Anhydrides of polyhydric alcohols. XIII. The amino derivatives of 1, 4:3, 6-dianhydromannitol, -sorbitol, and L-iditol and their behavior towards nitrous acidBashford, V. G.; Wiggins, L. F.Journal of the Chemical Society (1950), (), 371-4CODEN: JCSOA9; ISSN:0368-1769.cf. C.A. 43, 2939f; 44, 1O24d. 2, 5-Diamino-1, 4:3, 6-dianhydro-2, 5-didesoxysorbitol (I) (3.95 g.) in 25 cc. H2O, acidified with dil. HCl and treated with 4.5 g. NaNO2 in 25 cc. H2O, gives 1.27 g. 1,4:3,6-dianhydro-L-iditol (II), b0.05 115-30° (bath), [α]D 46.7° (Me2CO, c 1.9), characterized as the 2,5-bis(methylsulfonyl) deriv., m. 156.5° the small yield indicates that II was not the only product; the 2, 5-didesoxymannitol analog (III) of I also gives II; in the formation of II, deamination is accompanied by Walden inversion at C5 in the case of I and at both C2 and C5 in the case of III; 2,5-bis(p-tolylsulfonyl) deriv. (IV) of II, m. 90°, [α]D38.2°(CHCl3, c 2). IV(42g.) in 900 cc. MeOH, satd. at 0° with NH3, heated 30 hrs. at 160-70°, and the residue heated at 100° with 50 g. Ba(OH)2 in 400 cc. H2O, gives 2 g. 2,5-imino-1,4:3,6-dianhydro-2,5-didesoxy-D-mannitol (?) (V), m. 99-100°, [α]D 90.5° (CHCl3, c 2.65) [picrate, yellow, m. 219-20° (decompn.); HCl salt, m. 280-90° (decompn.); oxalate, m. 243°]; with NaNO2 in dil. AcOH, V forms an N-NO deriv., m. 121.5°, [α]D -323° (CHCl3, c 2.43). The conversion of I and III into II is paralleled by the epimerization of 2,4:3,5-dimethylene-D-manno- and -D-glucosaccharic acids (C.A. 38, 2631.8).
- 23Thiyagarajan, S.; Gootjes, L.; Vogelzang, W.; Wu, J.; van Haveren, J.; van Es, D. S. Chiral building blocks from biomass: 2,5-diamino-2,5-dideoxy-1,4-3,6-dianhydroiditol. Tetrahedron 2011, 67, 383– 389, DOI: 10.1016/j.tet.2010.11.031Google Scholar23https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXhsFCqs7rO&md5=dbb1b78c2a02975049665b0a4b5fca25Chiral building blocks from biomass: 2,5-diamino-2,5-dideoxy-1,4-3,6-dianhydroiditolThiyagarajan, Shanmugam; Gootjes, Linda; Vogelzang, Willem; Wu, Jing; van Haveren, Jacco; van Es, Daan S.Tetrahedron (2011), 67 (2), 383-389CODEN: TETRAB; ISSN:0040-4020. (Elsevier Ltd.)An efficient route towards the synthesis of 2,5-diamino-2,5-dideoxy-1,4-3,6-dianhydroiditol 4 has been developed resulting in significant improvements in both isolated yields and purity when compared to literature procedures. As a consequence, resin-grade 2,5-diamino-2,5-dideoxy-1,4-3,6-dianhydroiditol 4 has become available for lab. scale step-growth polymer synthesis. Addnl., an interesting renewable chiral 2-amino-2-deoxy-1,4-3,6-dianhydroiditol, has been isolated.
- 24Thiyagarajan, S.; Gootjes, L.; Vogelzang, W.; van Haveren, J.; Lutz, M.; van Es, D. S. Renewable rigid diamines: efficient, stereospecific synthesis of high purity isohexide diamines. ChemSusChem 2011, 4, 1823– 1829, DOI: 10.1002/cssc.201100398Google Scholar24https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXhsFeqt73O&md5=baeaf5ec2bbd64becd55551985f14c72Renewable Rigid Diamines: Efficient, Stereospecific Synthesis of High Purity Isohexide DiaminesThiyagarajan, Shanmugam; Gootjes, Linda; Vogelzang, Willem; van Haveren, Jacco; Lutz, Martin; van Es, Daan S.ChemSusChem (2011), 4 (12), 1823-1829CODEN: CHEMIZ; ISSN:1864-5631. (Wiley-VCH Verlag GmbH & Co. KGaA)We report an efficient three-step strategy for synthesizing rigid, chiral isohexide diamines derived from 1,4:3,6-dianhydrohexitols. These biobased chiral building blocks are presently the subject of several investigations (in our and several other groups) because of their application in high-performance biobased polymers, such as polyamides and polyurethanes. Among the three possible stereo-isomers, dideoxy-diamino isoidide and dideoxy-diamino isosorbide can be synthesized from isomannide and isosorbide resp. in high yield with abs. stereo control. Furthermore, by using this methodol. dideoxy-amino isomannide - a tricyclic adduct - was obtained starting from isoidide in high yield. Our improved synthetic route is a valuable advance towards meeting scale and purity demands for evaluating the properties of new biobased performance materials, which will benefit the development of these plastics.
- 25Kuszmann, J.; Medgyes, G. Synthesis and biological activity of 1,4:3,6-dianhydro-2,5-diazido-2,5-dideoxyhexitols. Carbohydr. Res. 1980, 85, 259– 269, DOI: 10.1016/s0008-6215(00)84675-3Google Scholar25https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL3MXht1WrsL8%253D&md5=f0a8154a0a54251bcce73388a2dcc70cSynthesis and biological activity of 1,4:3,6-dianhydro-2,5-diazido-2,5-dideoxyhexitolsKuszmann, Janos; Medgyes, GaborCarbohydrate Research (1980), 85 (2), 259-69CODEN: CRBRAT; ISSN:0008-6215.Reaction of 1,4:3,6-dianhydro-2,5-di-O-mesyl- and -tosyl-D-mannitol with NaN3 afforded the 2,5-diazido-L-iditol deriv. The analogous D-glucitol isomer was obtained in a similar reaction starting from the corresponding D-glucitol derivs., and showed significant, hypnotic activity (no data). For establishing the structure-activity relationship, 1,4:3,6-dianhydro-2,5-diazido-2,5-dideoxy-L-mannitol (I), as well as its antipode (II), was synthesized, starting from D-mannitol. I was as effective as Doriden (3-ethyl-3-phenylglutarimide), a well known hypnotic drug. II and the bioisosteric 1(4),3(6)-dithio deriv. were, however, inactive.
- 26Bähn, S.; Imm, S.; Neubert, L.; Zhang, M.; Neumann, H.; Beller, M. The catalytic amination of alcohols. ChemCatChem 2011, 3, 1853– 1864, DOI: 10.1002/cctc.201100255Google ScholarThere is no corresponding record for this reference.
- 27Pera-Titus, M.; Shi, F. Catalytic amination of biomass-based alcohols. ChemSusChem 2014, 7, 720– 722, DOI: 10.1002/cssc.201301095Google Scholar27https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXjtVGmsrY%253D&md5=bbb98a0b39689f5ecf960a64fa812d0bCatalytic Amination of Biomass-Based AlcoholsPera-Titus, Marc; Shi, FengChemSusChem (2014), 7 (3), 720-722CODEN: CHEMIZ; ISSN:1864-5631. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. Although alc. amination reactions were extensively studied since the early decades of the last century, the realization of sustainable development in the amine industry by implementing biomass-based alcs. as starting materials is still in its infancy. Catalytic systems based on Ru and Ir operating by means of the BH mechanism are currently available for performing such reactions. However, to reduce the prodn. costs, homogeneous systems based on cheaper metals operating by nucleophilic substitution, as well as supported metal nanoparticles (Ni, Co, Cu, Pd, Au) on low- alk. supports, are highly desired and new developments are expected to occur soon. With these catalysts in hand, new research areas for amination involving biomass-based alcs. with market opportunities are expected to be developed during this decade.
- 28Pingen, D.; Diebolt, O.; Vogt, D. Direct amination of bio-alcohols using ammonia. ChemCatChem 2013, 5, 2905– 2912, DOI: 10.1002/cctc.201300407Google Scholar28https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXhtVGktbfJ&md5=ad5a8a6a4ed29046a42e7eb4bfe6dff0Direct Amination of Bio-Alcohols Using AmmoniaPingen, Dennis; Diebolt, Olivier; Vogt, DieterChemCatChem (2013), 5 (10), 2905-2912CODEN: CHEMK3; ISSN:1867-3880. (Wiley-VCH Verlag GmbH & Co. KGaA)A slightly adapted catalyst system has been successfully applied in the direct amination of primary and secondary alcs. Moreover, the applicability to diols has been shown, giving high selectivity towards the primary diamines. It was found that the Ru/P ratio as well as the amt. of ammonia used are highly important in this system, esp. for higher substrate loadings. The catalyst was employed on a larger batch scale for the conversion of isomannide to the corresponding diamine. Addnl., it was shown that the catalyst is stable for at least six consecutive runs. No significant loss of activity and selectivity was obsd.
- 29Wright, L. W.; Brandner, J. D. Catalytic Isomerization of Polyhydric Alcohols.1 II. The Isomerization of Isosorbide to Isomannide and Isoidide. J. Org. Chem. 1964, 29, 2979– 2982, DOI: 10.1021/jo01033a043Google Scholar29https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaF2cXkvVCmtr0%253D&md5=1a81c31e7a0756aabc537182147720b1Catalytic isomerization of polyhydric alcohols. II. The isomerization of isosorbide to isomannide and isoidideWright, L. W.; Brandner, J. D.Journal of Organic Chemistry (1964), 29 (10), 2979-82CODEN: JOCEAH; ISSN:0022-3263.At 220-240° and 150 atm. H pressure the reversible interconversion of the 1,4:3,6-dianhydrohexitols of D-glucitol, D-mannitol, and L-iditol reaches a steady state after 2-6 h. in the presence of Ni-kieselguhr catalyst. At this time the approx. concns. are 57 % 1,4:3,6-dianhydro-L-iditol, 36% 1,4:3,6-dianhydro-D-glucitol; and 7% 1,4:3,6-dianhydro-D-mannitol. These figures are shown to be consistent with probability considerations, i.e., the relative amts. of the dianhydrohexitols are related to the probability of a given OH group being either exo or endo. Taking the steady-state mole fraction of 1,4:3,6-dianhydro-L-iditol as 0.57, it is calcd. that the probability of a OH being exo is 3 times the probability of its being endo. Calcn. of the mole fraction of the other 2 anhydrohexitols on the basis of these relative probabilities yields values in close agreement with those found exptl. The isomerization is strongly accelerated by increasing alky. of the catalyst-dianhydrohexitol slurry. Cf. CA 56, 11677c.
- 30Brandner, D.; Wright, L. W. Process for producing isoidide. U.S. Patent no. 3,023,223 A, 1962.Google ScholarThere is no corresponding record for this reference.
- 31Schelwies, M.; Brinks, M.; Schaub, T.; Melder, J.-P.; Paciello, R.; Merger, M. Process for the homogeneously catalyzed amination of alcohols with ammonia in the presence of a complex catalyst which comprises nonanionic ligands. Patent no. WO 2014016241 A1 2014Google ScholarThere is no corresponding record for this reference.
- 32Bahé, F.; Grand, L.; Cartier, E.; Jacolot, M.; Moebs-Sanchez, S.; Portinha, D.; Fleury, E.; Popowycz, F. Direct amination of isohexides via borrowing hydrogen methodology: regio- and stereoselective issues. Eur. J. Org. Chem. 2020, 2020, 599– 608, DOI: 10.1002/ejoc.201901661Google Scholar32https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXhs1ertr8%253D&md5=305cec7a4a4805b400ea26905f0549c3Direct Amination of Isohexides via Borrowing Hydrogen Methodology: Regio- and Stereoselective IssuesBahe, Florian; Grand, Lucie; Cartier, Elise; Jacolot, Maiwenn; Moebs-Sanchez, Sylvie; Portinha, Daniel; Fleury, Etienne; Popowycz, FlorenceEuropean Journal of Organic Chemistry (2020), 2020 (5), 599-608CODEN: EJOCFK; ISSN:1099-0690. (Wiley-VCH Verlag GmbH & Co. KGaA)The regio and diastereoselective direct mono or diamination of bio-based isohexides (isosorbide and isomannide) has been developed through borrowing hydrogen (BH) methodol. using a cooperative catalysis between an iridium complex and a Bronsted acid. The access to chiral amino-alc. (NH2-OH) and diamine (NH2-NH2), interesting optically pure bio-based monomers, was also proposed using BH strategy.
- 33Pfützenreuter, R.; Rose, M. Aqueous-phase amination of biogenic isohexides by using Ru/C as a solid catalyst. ChemCatChem 2016, 8, 251– 255, DOI: 10.1002/cctc.201501077Google Scholar33https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXhvFenurfE&md5=0207c7b9b607c92581601cfc22cfa8d3Aqueous-Phase Amination of Biogenic Isohexides by using Ru/C as a Solid CatalystPfuetzenreuter, Rebecca; Rose, MarcusChemCatChem (2016), 8 (1), 251-255CODEN: CHEMK3; ISSN:1867-3880. (Wiley-VCH Verlag GmbH & Co. KGaA)An aq. phase process for the amination of biogenic isohexides with ammonia using Ru/C as solid catalyst is reported. The products possess great potential for the prodn. of biogenic polyamides. From the diol substrates amino alcs. and diamines are derived as main products with significantly varying product selectivity depending on the substrates' stereochem. The catalytic reaction is performed at temps. of 140-180 °C. The Ru/C catalyst shows a high activity in the aq. phase, which is beneficial for the transformation of the polar biogenic substrates. Despite the basic conditions, the metal leaching is negligible and the catalyst can be recycled easily in batch operation and, thus, enables further process development due to the mild and scalable reaction conditions.
- 34Niemeier, J.; Engel, R. V.; Rose, M. Is water a suitable solvent for the catalytic amination of alcohols?. Green Chem. 2017, 19, 2839– 2845, DOI: 10.1039/c7gc00422bGoogle Scholar34https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXkvF2ku7k%253D&md5=ecdabfe39aabca9726addeebf734d262Is water a suitable solvent for the catalytic amination of alcohols?Niemeier, Johannes; Engel, Rebecca V.; Rose, MarcusGreen Chemistry (2017), 19 (12), 2839-2845CODEN: GRCHFJ; ISSN:1463-9262. (Royal Society of Chemistry)The catalytic conversion of biomass and biogenic platform chems. typically requires the use of solvents. Water is present already in the raw materials and in most cases a suitable solvent for the typically highly polar substrates. Hence, the development of novel catalytic routes for further processing would profit from the optimization of the reaction conditions in the aq. phase mainly for energetic reasons by avoiding the initial water sepn. Herein, we report the amination of biogenic alcs. in aq. solns. using solid Ru-based catalysts and ammonia as a reactant. The influence of different support materials and bimetallic catalysts is investigated for the amination of isomannide as a biogenic diol. Most importantly, the transferability of the reaction conditions to various other primary and secondary alcs. is successfully proved. Hence, water appears to be a suitable solvent for the sustainable prodn. of biogenic amines and offers great potential for further process development.
- 35Hu, H.; Ramzan, A.; Wischert, R.; Jerôme, F.; Michel, C.; de Olivera Vigier, K.; Pera-Titus, M. Pivotal role of H2 in the isomerisation of isosorbide over a Ru/C catalyst. Catal. Sci. Technol. 2021, 11, 7973– 7981, DOI: 10.1039/d1cy01709hGoogle Scholar35https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3MXit1KmtLjF&md5=2c5f64a0017eb471c2b29a740e4d49a7Pivotal role of H2 in the isomerization of isosorbide over a Ru/C catalystHu, H.; Ramzan, A.; Wischert, R.; Jerome, F.; Michel, C.; de Olivera Vigier, K.; Pera-Titus, M.Catalysis Science & Technology (2021), 11 (24), 7973-7981CODEN: CSTAGD; ISSN:2044-4753. (Royal Society of Chemistry)Isosorbide isomerisation is a known reaction that can proceed over Ru and Ni-based heterogeneous catalysts. As a rule, an exogenous H2 pressure (40-100 bar) is required, even though H2 does not participate stoichiometrically in the reaction. By marrying expts. with DFT computations, we ascribe the role of H2 in isosorbide isomerisation to a coverage effect on the catalyst surface. We demonstrate the possibility of conducting the reaction at a low H2 pressure either in the presence of an inert gas to increase H2 soly. in an underlying solvent or using 2-propanol as a hydrogen donor. This might benefit the economy and safety of a potential industrial process.
- 36Hausoul, P. J. C.; Negahdar, L.; Schute, K.; Palkovits, R. Unravelling the Ru-catalyzed hydrogenolysis of biomass-based polyols under neutral and acidic conditions. ChemSusChem 2015, 8, 3323– 3330, DOI: 10.1002/cssc.201500493Google Scholar36https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXhsVOms7fO&md5=f641f25935c7fa50f1829a2edf002e16Unravelling the Ru-Catalyzed Hydrogenolysis of Biomass-Based Polyols under Neutral and Acidic ConditionsHausoul, Peter J. C.; Negahdar, Leila; Schute, Kai; Palkovits, ReginaChemSusChem (2015), 8 (19), 3323-3330CODEN: CHEMIZ; ISSN:1864-5631. (Wiley-VCH Verlag GmbH & Co. KGaA)The aq. Ru/C-catalyzed hydrogenolysis of biomass-based polyols such as erythritol, xylitol, sorbitol, and cellobitol is studied under neutral and acidic conditions. For the first time, the complete product spectrum of C2-C6 polyols is identified and, based on a thorough anal. of the reaction mixts., a comprehensive reaction mechanism is proposed, which consists of (de)hydrogenation, epimerization, decarbonylation, and deoxygenation reactions. The data reveal that the Ru-catalyzed deoxygenation reaction is highly selective for the cleavage of terminal hydroxyl groups. Changing from neutral to acidic conditions suppresses decarbonylation, consequently increasing the selectivity towards deoxygenation.
- 37Wang, T.; Ibañez, J.; Wang, K.; Fang, L.; Sabbe, M.; Michel, C.; Paul, S.; Pera-Titus, M.; Sautet, P. Rational design of selective metal catalysts for alcohol amination with ammonia. Nat. Catal. 2019, 2, 773– 779, DOI: 10.1038/s41929-019-0327-2Google Scholar37https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXhsFWlt7nF&md5=a01eb8c81f84ae1159a04fe6225f7d56Rational design of selective metal catalysts for alcohol amination with ammoniaWang, Tao; Ibanez, Javier; Wang, Kang; Fang, Lin; Sabbe, Maarten; Michel, Carine; Paul, Sebastien; Pera-Titus, Marc; Sautet, PhilippeNature Catalysis (2019), 2 (9), 773-779CODEN: NCAACP; ISSN:2520-1158. (Nature Research)The lack of selectivity for the direct amination of alcs. with ammonia (a modern and clean route for the synthesis of primary amines) is an unsolved problem. Here, we combine first-principles calcns., scaling relations, kinetic simulations and catalysis expts. to det. the key factors that govern the activity and selectivity of metal catalysts for this reaction. We show that the loss of selectivity towards primary amines is linked to a surface-mediated C-N bond coupling between two N-contg. intermediates: CH3NH and CH2NH. The barrier for this step is low enough to compete with the main surface hydrogenation reactions and it can be used as a descriptor for selectivity. The activity and selectivity maps (using the C and O adsorption energies as descriptors) were combined for the computational screening of 348 dil. bimetallic catalysts. Among the best theor. candidates, Co98.5Ag1.5 and Co98.5Ru1.5 (5 wt% Co) were identified exptl. to be the most promising catalysts.
- 38Honkala, K.; Hellman, A.; Remediakis, I. N.; Logadottir, A.; Carlsson, A.; Dahl, S.; Christensen, C. H.; Nørskov, J. K. Ammonia synthesis from first-principles calculations. Science 2005, 307, 555– 558, DOI: 10.1126/science.1106435Google Scholar38https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2MXmslOjuw%253D%253D&md5=4c6066f84d200c9c587b6dd1dcd3b0bfAmmonia Synthesis from First-Principles CalculationsHonkala, K.; Hellman, A.; Remediakis, I. N.; Logadottir, A.; Carlsson, A.; Dahl, S.; Christensen, C. H.; Norskov, J. K.Science (Washington, DC, United States) (2005), 307 (5709), 555-558CODEN: SCIEAS; ISSN:0036-8075. (American Association for the Advancement of Science)The rate of ammonia synthesis over a nanoparticle ruthenium catalyst can be calcd. directly on the basis of a quantum chem. treatment of the problem using d. functional theory. We compared the results to measured rates over a ruthenium catalyst supported on magnesium aluminum spinel. When the size distribution of ruthenium particles measured by transmission electron microscopy was used as the link between the catalyst material and the theor. treatment, the calcd. rate was within a factor of 3 to 20 of the exptl. rate. This offers hope for computer-based methods in the search for catalysts.
- 39Lu, X.; Zhang, J.; Chen, W.-K.; Roldan, A. Kinetic and mechanistic analysis of NH3 decomposition on Ru(0001), Ru(111) and Ir(111) surfaces. Nanoscale Adv 2021, 3, 1624– 1632, DOI: 10.1039/d1na00015bGoogle Scholar39https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3MXjsVOgurg%253D&md5=e3fd50b5f5239cd7fe9e9fc3f182ed4bKinetic and mechanistic analysis of NH3 decomposition on Ru(0001), Ru(111) and Ir(111) surfacesLu, Xiuyuan; Zhang, Jing; Chen, Wen-Kai; Roldan, AlbertoNanoscale Advances (2021), 3 (6), 1624-1632CODEN: NAADAI; ISSN:2516-0230. (Royal Society of Chemistry)We investigated the catalytic NH3 decompn. on Ru and Ir metal surfaces using d. functional theory. The reaction mechanisms were unraveled on both metals, considering that, on the nano-scale, Ru particles may also present an fcc structure, hence, leading to three energy profiles. We implemented thermodn. and kinetic parameters obtained from DFT into microkinetic simulations. Batch reactor simulations suggest that hydrogen generation starts at 400 K, 425 K and 600 K on Ru(111), Ru(0001) and Ir(111) surfaces, resp., in excellent agreement with expts. During the reaction, the main surface species on Ru are NH, N and H, whereas on Ir(111), it is mainly NH. The rate-detg. step for all surfaces is the formation of mol. nitrogen. We also performed temp.-programmed reaction simulations and inspected the desorption spectra of N2 and H2 as a function of temp., which highlighted the importance of N coverage on the desorption rate.
- 40Benndorf, C.; Madey, T. E. Adsorption and orientation of NH3 on Ru(001). Surf. Sci. 1983, 135, 164– 183, DOI: 10.1016/0039-6028(83)90217-0Google Scholar40https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL2cXnsVGisw%253D%253D&md5=f07ea305e45db23b65297650120d78d3Adsorption and orientation of ammonia on ruthenium(001)Benndorf, Carsten; Madey, Theodore E.Surface Science (1983), 135 (1-3), 164-83CODEN: SUSCAS; ISSN:0039-6028.The interaction of NH3 with clean Ru(001) surfaces was studied by using LEED ESDIAD (electron stimulated desorption ion angular distribution), TDS (thermal desorption spectroscopy), and work function changes (Δ.vphi.). Four different binding states (denoted as α1, α2, β, and γ) were detected with TDS. At low coverages, NH3 desorbs from the α1 state with a TDS peak max. at ∼310 K. The broadening of the TDS peaks and their shift to lower temp. with increasing NH3 coverage are related to repulsive lateral interactions between neighboring NH3 mols. At higher NH3 coverages (θNH3 ⪆ 0.15), a 2nd desorption peak (α2) develops at 180 K, accompanied by a (2 × 2) LEED structure. With further increase of NH3 exposure a sharp desorption peak (β state) is found at 140 K, an is interpreted as due to NH3 species desorbing from a 2nd adsorption layer. Finally a desorption peak due to multilayer adsorption (γ state) is found at 115 K. At low NH3 coverages (α1 state), a halo-like H+ ESDIAD pattern gives evidence of randomly oriented or freely rotating NH3 monomers, bounded via the N atoms to the surface with the H atoms pointing away from the surface. This orientation of NH3 is supported by work function measurements showing a linear decrease of Δ.vphi. in the α1 state. Structural information concerning the adsorption geometry of NH3 in the β state was obtained from LEED and ESDIAD. During the formation of the 2nd NH3 layer (β) a (2√3 × 2√3)R30° LEED pattern is obsd. and is accompanied by an ESDIAD pattern with a hexagonal outline. A structural model of the β-state bonding, in which 2nd layer NH3 mols. are bonded via 3-fold H bonds to the 1st layer NH3, is proposed.
- 41Carabineiro, S. A. C.; Matveev, A. V.; Gorodetskii, V. V.; Nieuwenhuys, B. E. Selective oxidation of ammonia over Ru(0001). Surf. Sci. 2004, 555, 83– 93, DOI: 10.1016/j.susc.2004.02.022Google Scholar41https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXislyltrs%253D&md5=e6e8139a87dd657fb1c6d2a6c54759d5Selective oxidation of ammonia over Ru(0 0 0 1)Carabineiro, S. A. C.; Matveev, A. V.; Gorodetskii, V. V.; Nieuwenhuys, B. E.Surface Science (2004), 555 (1-3), 83-93CODEN: SUSCAS; ISSN:0039-6028. (Elsevier Science B.V.)The decompn. and oxidn. of NH3 were studied on the Ru(0 0 0 1) surface in the temp. range from 150 up to 800 K. The results were compared to those found for Ir(1 1 0) and Ir(5 1 0). TDS results showed that most of the NH3 is dissociatively adsorbed between 150 and 300 K, with formation of H2 around 300 K and N2 between 600 and 800 K. N2 desorption shifts to lower temps. with increasing surface O coverage. The products of NH3 oxidn. obsd. were N2, H2O, and N2O. Formation of NO was not found. Inhibition of the reaction presumably by N species was obsd. until 450 and 670 K, depending on the NH3/O2 ratios. Above those temps. the reaction started as manifested by a decrease in the NH3 and O2 pressures and a simultaneous increase in the H2O, N2 and N2O pressures.
- 42Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A., Jr; Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, J. M.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, O. ̈.; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian 09; Gaussian, Inc.: Wallingford, CT, 2016.Google ScholarThere is no corresponding record for this reference.
- 43Zhao, Y.; Truhlar, D. G. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: Two new functionals and systematic testing of four M06-class functionals and 12 other functionals. Theor. Chem. Acc. 2008, 120, 215– 241, DOI: 10.1007/s00214-007-0310-xGoogle Scholar43https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXltFyltbY%253D&md5=c31d6f319d7c7a45aa9b716220e4a422The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: two new functionals and systematic testing of four M06-class functionals and 12 other functionalsZhao, Yan; Truhlar, Donald G.Theoretical Chemistry Accounts (2008), 120 (1-3), 215-241CODEN: TCACFW; ISSN:1432-881X. (Springer GmbH)We present two new hybrid meta exchange-correlation functionals, called M06 and M06-2X. The M06 functional is parametrized including both transition metals and nonmetals, whereas the M06-2X functional is a high-nonlocality functional with double the amt. of nonlocal exchange (2X), and it is parametrized only for nonmetals. The functionals, along with the previously published M06-L local functional and the M06-HF full-Hartree-Fock functionals, constitute the M06 suite of complementary functionals. We assess these four functionals by comparing their performance to that of 12 other functionals and Hartree-Fock theory for 403 energetic data in 29 diverse databases, including ten databases for thermochem., four databases for kinetics, eight databases for noncovalent interactions, three databases for transition metal bonding, one database for metal atom excitation energies, and three databases for mol. excitation energies. We also illustrate the performance of these 17 methods for three databases contg. 40 bond lengths and for databases contg. 38 vibrational frequencies and 15 vibrational zero point energies. We recommend the M06-2X functional for applications involving main-group thermochem., kinetics, noncovalent interactions, and electronic excitation energies to valence and Rydberg states. We recommend the M06 functional for application in organometallic and inorganometallic chem. and for noncovalent interactions.
- 44Weigend, F.; Ahlrichs, R. Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys. Chem. Chem. Phys. 2005, 7, 3297– 3305, DOI: 10.1039/b508541aGoogle Scholar44https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2MXpsFWgu7o%253D&md5=a820fb6055c993b50c405ba0fc62b194Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracyWeigend, Florian; Ahlrichs, ReinhartPhysical Chemistry Chemical Physics (2005), 7 (18), 3297-3305CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)Gaussian basis sets of quadruple zeta valence quality for Rb-Rn are presented, as well as bases of split valence and triple zeta valence quality for H-Rn. The latter were obtained by (partly) modifying bases developed previously. A large set of more than 300 mols. representing (nearly) all elements-except lanthanides-in their common oxidn. states was used to assess the quality of the bases all across the periodic table. Quantities investigated were atomization energies, dipole moments and structure parameters for Hartree-Fock, d. functional theory and correlated methods, for which we had chosen Moller-Plesset perturbation theory as an example. Finally recommendations are given which type of basis set is used best for a certain level of theory and a desired quality of results.
- 45Weigend, F.; Ahlrichs, R. Accurate Coulomb-fitting basis sets for H to Rn. Phys. Chem. Chem. Phys. 2006, 8, 1057– 1065, DOI: 10.1039/b515623hGoogle Scholar45https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28Xhs12ntrc%253D&md5=314690393f1e21096541a317a80e563cAccurate Coulomb-fitting basis sets for H to RnWeigend, FlorianPhysical Chemistry Chemical Physics (2006), 8 (9), 1057-1065CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)A series of auxiliary basis sets to fit Coulomb potentials for the elements H to Rn (except lanthanides) is presented. For each element only one auxiliary basis set is needed to approx. Coulomb energies in conjunction with orbital basis sets of split valence, triple zeta valence and quadruple zeta valence quality with errors of typically below ca. 0.15 kJ mol-1 per atom; this was demonstrated in conjunction with the recently developed orbital basis sets of types def2-SV(P), def2-TZVP and def2-QZVPP for a large set of small mols. representing (nearly) each element in all of its common oxidn. states. These auxiliary bases are slightly more than three times larger than orbital bases of split valence quality. Compared to non-approximated treatments, computation times for the Coulomb part are reduced by a factor of ca. 8 for def2-SV(P) orbital bases, ca. 25 for def2-TZVP and ca. 100 for def2-QZVPP orbital bases.
- 46Luchini, G.; Alegre-Requena, J. V.; Funes-Ardoiz, I.; Paton, R. S. GoodVibes: automated thermochemistry for heterogeneous computational chemistry data; F1000 Research Ltd, 2020, Vol 9.Google ScholarThere is no corresponding record for this reference.
- 47Alecu, I. M.; Zheng, J.; Zhao, Y.; Truhlar, D. G. Computational thermochemistry: scale factor databases and scale factors for vibrational frequencies obtained from electronic model chemistries. J. Chem. Theory Comput. 2010, 6, 2872– 2887, DOI: 10.1021/ct100326hGoogle Scholar47https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXhtVGrsb7E&md5=f0104fdf97972d89e7cb60e634b19c4bComputational Thermochemistry: Scale Factor Databases and Scale Factors for Vibrational Frequencies Obtained from Electronic Model ChemistriesAlecu, I. M.; Zheng, Jingjing; Zhao, Yan; Truhlar, Donald G.Journal of Chemical Theory and Computation (2010), 6 (9), 2872-2887CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Optimized scale factors for calcg. vibrational harmonic and fundamental frequencies and zero-point energies have been detd. for 145 electronic model chemistries, including 119 based on approx. functionals depending on occupied orbitals, 19 based on single-level wave function theory, three based on the neglect-of-diat.-differential-overlap, two based on doubly hybrid d. functional theory, and two based on multicoefficient correlation methods. Forty of the scale factors are obtained from large databases, which are also used to derive two universal scale factor ratios that can be used to interconvert between scale factors optimized for various properties, enabling the derivation of three key scale factors at the effort of optimizing only one of them. A reduced scale factor optimization model is formulated in order to further reduce the cost of optimizing scale factors, and the reduced model is illustrated by using it to obtain 105 addnl. scale factors. Using root-mean-square errors from the values in the large databases, we find that scaling reduces errors in zero-point energies by a factor of 2.3 and errors in fundamental vibrational frequencies by a factor of 3.0, but it reduces errors in harmonic vibrational frequencies by only a factor of 1.3. It is shown that, upon scaling, the balanced multicoefficient correlation method based on coupled cluster theory with single and double excitations (BMC-CCSD) can lead to very accurate predictions of vibrational frequencies. With a polarized, minimally augmented basis set, the d. functionals with zero-point energy scale factors closest to unity are MPWLYP1M (1.009), τHCTHhyb (0.989), BB95 (1.012), BLYP (1.013), BP86 (1.014), B3LYP (0.986), MPW3LYP (0.986), and VSXC (0.986).
- 48Ribeiro, R. F.; Marenich, A. V.; Cramer, C. J.; Truhlar, D. G. Use of solution-phase vibrational frequencies in continuum models for the free energy of solvation. J. Phys. Chem. B 2011, 115, 14556– 14562, DOI: 10.1021/jp205508zGoogle Scholar48https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXhsFSjtr3O&md5=3a164fbab7255d92e1099064e7f72261Use of Solution-Phase Vibrational Frequencies in Continuum Models for the Free Energy of SolvationRibeiro, Raphael F.; Marenich, Aleksandr V.; Cramer, Christopher J.; Truhlar, Donald G.Journal of Physical Chemistry B (2011), 115 (49), 14556-14562CODEN: JPCBFK; ISSN:1520-5207. (American Chemical Society)We find that vibrational contributions to a solute's free energy are in general insensitive to whether the solute vibrational frequencies are computed in the gas phase or in soln. In most cases, the difference is smaller than the intrinsic error in solvation free energies assocd. with the continuum approxn. to solvation modeling, although care must be taken to avoid spurious results assocd. with limitations in the quantum-mech. harmonic-oscillator approxn. for very low-frequency mol. vibrations. We compute solute vibrational partition functions in aq. and carbon tetrachloride soln. and compare them to gas-phase mol. partition functions computed with the same level of theory and the same quasiharmonic approxn. for the diverse and extensive set of mols. and ions included in the training set of the SMD continuum solvation model, and we find mean unsigned differences in vibrational contributions to the solute free energy of only about 0.2 kcal/mol. On the basis of these results and a review of the theory, we conclude, in contrast to previous work, that using partition functions computed for mols. optimized in soln. is a correct and useful approach for averaging over solute degrees of freedom when computing free energies of solutes in soln., and it is moreover recommended for cases where liq. and gas-phase solute structures differ appreciably or when stationary points present in liq. soln. do not exist in the gas phase, for which we provide some examples. When gas-phase and soln.-phase geometries and frequencies are similar, the use of gas-phase geometries and frequencies is a useful approxn.
- 49Li, Y.-P.; Gomes, J.; Mallikarjun Sharada, S.; Bell, A. T.; Head-Gordon, M. Improved force-field parameters for QM/MM simulations of the energies of adsorption for molecules in zeolites and a free rotor correction to the rigid rotor harmonic oscillator model for adsorption enthalpies. J. Phys. Chem. C 2015, 119, 1840– 1850, DOI: 10.1021/jp509921rGoogle Scholar49https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXitFehsbbJ&md5=d5fd7f41d9a44592826c59eea27f73c1Improved Force-Field Parameters for QM/MM Simulations of the Energies of Adsorption for Molecules in Zeolites and a Free Rotor Correction to the Rigid Rotor Harmonic Oscillator Model for Adsorption EnthalpiesLi, Yi-Pei; Gomes, Joseph; Mallikarjun Sharada, Shaama; Bell, Alexis T.; Head-Gordon, MartinJournal of Physical Chemistry C (2015), 119 (4), 1840-1850CODEN: JPCCCK; ISSN:1932-7447. (American Chemical Society)Quantum mechanics/mol. mechanics (QM/MM) simulations provide an efficient avenue for studying reactions catalyzed in zeolite systems; however, the accuracy of such calcns. is highly dependent on the zeolite MM parameters used. Previously reported parameters (P1), which were chosen to minimize the root mean square (RMS) deviations of adsorption energies compared with full QM ωB97X-D/6-31+G** adsorption energies, are shown to overestimate binding energies compared with exptl. values, particularly for larger substrates. To address this issue, a new parameter set (P2) is derived by rescaling the previously reported characteristic energies of the Lennard-Jones potential in P1. The accuracy of the thermal correction for adsorption enthalpies detd. by the rigid rotor-harmonic oscillator approxn. (RRHO) is examd. and shown to be improved by treating low-lying vibrational modes as free translational and rotational modes via a quasi-RRHO model. With P2 and quasi-RRHO, adsorption energies calcd. with QM/MM agree with exptl. values with an RMS error of 1.8 kcal/mol for both nonpolar and polar mols. adsorbed in MFI, H-MFI, and H-BEA. By contrast, the RMS error for the same test sets obtained using parameter set P1 is 8.3 kcal/mol. Glucose-fructose isomerization catalyzed by Sn-BEA is taken as an example to demonstrate that improved values for apparent activation energies can be obtained using the methodol. reported here. With parameter set P2, the apparent activation energy calcd. with QM/MM reproduces the exptl. value to within 1 kcal/mol. By contrast, using parameter set P1, the error is -12.9 kcal/mol.
- 50Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. J. Phys. Rev. B 1996, 54, 11169– 11186, DOI: 10.1103/physrevb.54.11169Google Scholar50https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK28Xms1Whu7Y%253D&md5=9c8f6f298fe5ffe37c2589d3f970a697Efficient 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.
- 51Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865– 3868, DOI: 10.1103/physrevlett.77.3865Google Scholar51https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK28XmsVCgsbs%253D&md5=55943538406ee74f93aabdf882cd4630Generalized 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.
- 52Steinmann, S.; Corminboeuf, C. A generalized-gradient approximation exchange hole model for dispersion coefficients. J. Chem. Phys. 2011, 134, 044117, DOI: 10.1063/1.3545985Google Scholar52https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXht1Giur4%253D&md5=c28723818938f9c1326bcb551835d562A generalized-gradient approximation exchange hole model for dispersion coefficientsSteinmann, Stephan N.; Corminboeuf, ClemenceJournal of Chemical Physics (2011), 134 (4), 044117/1-044117/5CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)A simple method for computing accurate d.-dependent dispersion coeffs. is presented. The dispersion coeffs. are modeled by a generalized gradient-type approxn. to Becke and Johnson's exchange hole dipole moment formalism. Our most cost-effective variant, based on a disjoint description of atoms in a mol., gives mean abs. errors in the C6 coeffs. for 90 complexes below 10%. The inclusion of the missing long-range van der Waals interactions in d. functionals using the derived coeffs. in a pair wise correction leads to highly accurate typical noncovalent interaction energies. (c) 2011 American Institute of Physics.
- 53Blöchl, P. E. Projector augmented-wave method. Phys. Rev. B 1994, 50, 17953– 17979, DOI: 10.1103/physrevb.50.17953Google Scholar53https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC2sfjslSntA%253D%253D&md5=1853d67af808af2edab58beaab5d3051Projector augmented-wave methodBlochlPhysical review. B, Condensed matter (1994), 50 (24), 17953-17979 ISSN:0163-1829.There is no expanded citation for this reference.
- 54Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758– 1775, DOI: 10.1103/physrevb.59.1758Google Scholar54https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXkt12nug%253D%253D&md5=78a73e92a93f995982fc481715729b14From 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.
- 55Hermes, E. D.; Janes, A. N.; Schmidt, J. R. M. Micki: A python-based object-oriented microkinetic modeling code. J. Chem. Phys. 2019, 151, 014112, DOI: 10.1063/1.5109116Google Scholar55https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXhtlahs7nK&md5=97a4449d51c0d1a40a261e77ea73f76aMicki: A python-based object-oriented microkinetic modeling codeHermes, Eric D.; Janes, Aurora N.; Schmidt, J. R.Journal of Chemical Physics (2019), 151 (1), 014112/1-014112/13CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)We have developed a flexible, general-purpose microkinetic modeling code, Micki, to analyze complex, heterogeneously catalyzed chem. reactions based upon first-principles calcns. This Python-based code is modular and object oriented, framing the development of microkinetic models in familiar chem. terms. We also present novel approaches, incorporated into Micki, to describe diffusion limited reactions, multidentate bindings, thermodynamically consistent lateral interactions, and Bronsted-Evans-Polanyi ests. of changes in barrier heights. Micki has built-in modules for subsequent anal. of microkinetic models, including degree of rate control and rate order. As a demonstration of the power and flexibility of the code, we build a microkinetic model for the water-gas shift reaction and compare to previously published exptl. results and microkinetic models, showing that Micki can quant. reproduce exptl. turnover frequencies with minimal empirical optimization. (c) 2019 American Institute of Physics.
Cited By
This article has not yet been cited by other publications.
Article Views
Altmetric
Citations
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
References
This article references 55 other publications.
- 1Wittcoff, H. A.; Reuben, B. G.; Plotkin, J. S. Industrial Organic Chemicals. 2nd ed., Wiley, NY, 2004.There is no corresponding record for this reference.
- 2Lawrence, S. A. Amines: Synthesis, Properties and Applications; Cambridge University Press, 2004.There is no corresponding record for this reference.
- 3Roose, P.; Eller, K.; Henkes, E.; Rossbacher, R.; Höke, H. Amines, Aliphatic, Ullmann’s Encyclopedia of Industrial Chemistry; Wiley-VCH: Weinheim, 2015, p 1.There is no corresponding record for this reference.
- 4Imm, S.; Bahn, S.; Zhang, M.; Neubert, L.; Neumann, H.; Klasovsky, F.; Pfeffer, J.; Haas, T.; Beller, M. Improved ruthenium-catalyzed amination of alcohols with ammonia: synthesis of diamines and amino esters. Angew. Chem., Int. Ed. 2011, 50, 7599– 7603, DOI: 10.1002/anie.2011031994https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXosFemtL4%253D&md5=f5cdc49c0486d53561cb1b24a47660beImproved ruthenium-catalyzed amination of alcohols with ammonia: synthesis of diamines and amino estersImm, Sebastian; Baehn, Sebastian; Zhang, Min; Neubert, Lorenz; Neumann, Helfried; Klasovsky, Florian; Pfeffer, Jan; Haas, Thomas; Beller, MatthiasAngewandte Chemie, International Edition (2011), 50 (33), 7599-7603, S7599/1-S7599/11CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)A homogeneous selective catalytic amination of primary and secondary diols with ammonia is described. Catalyzed by [Ru(CO)ClH(PPh3)3]/Xantphos, the corresponding diamines were obtained in good yields using this method.
- 5Delidovich, I.; Hausoul, C.; Deng, L.; Pfützenreuter, R.; Rose, M.; Palkovits, R. Alternative Monomers Based on Lignocellulose and Their Use for Polymer Production. Chem. Rev. 2016, 116, 1540– 1599, DOI: 10.1021/acs.chemrev.5b003545https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXhslGmsrbN&md5=202111f6f0d0f0517fc592e5e4c24f74Alternative Monomers Based on Lignocellulose and Their Use for Polymer ProductionDelidovich, Irina; Hausoul, Peter J. C.; Deng, Li; Pfuetzenreuter, Rebecca; Rose, Marcus; Palkovits, ReginaChemical Reviews (Washington, DC, United States) (2016), 116 (3), 1540-1599CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review. In recent years, the valorization of lignocellulose as feedstock for the prodn. of platform chems. has found tremendous attention. At the same time, the further transformation of these platform chems. into alternative monomers has been an area of growing research interest. Nevertheless, one has to point out that several of the monomers discussed in this review have already been synthesized by means of org. synthesis several decades ago. However, the renewed interest in lignocellulose based monomers entailed the design of green catalytic synthesis protocols to access these monomers.
- 6Pelckmans, M.; Renders, T.; Van de Vyver, S.; Sels, B. F. Bio-based amines through sustainable heterogeneous catalysis. Green Chem. 2017, 19, 5303– 5331, DOI: 10.1039/c7gc02299a6https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXhsFWiu77J&md5=cd61d96d3dd6031bfb81ead1ceb815c0Bio-based amines through sustainable heterogeneous catalysisPelckmans, M.; Renders, T.; Van de Vyver, S.; Sels, B. F.Green Chemistry (2017), 19 (22), 5303-5331CODEN: GRCHFJ; ISSN:1463-9262. (Royal Society of Chemistry)The prodn. of amines from biomass is a growing field of interest. Particularly the amination of bio-based alcs. receives a lot of attention. In this review, we discuss recent progress in the development of efficient heterogeneous catalysts. The substrate scope for the prodn. of bio-based amines is not limited to (hemi)cellulosic alcs. Other platform chems. that originate from different biomass fractions, such as lignin, oils, chitin and protein, are also suitable feedstock for the prodn. of amines. This comprehensive review first provides an overview of the available bio-based feedstock candidates. The following section is devoted to the sustainable reaction routes that are available to carry out the desired amination reactions. Next, state-of-the-art technologies are summarized for each substrate class, focussing on heterogeneous catalysis. Special attention is dedicated to the sustainability of the discussed reaction routes. Finally, a crit. discussion is provided, together with current challenges and future perspectives regarding the industrial prodn. of bio-based amine chems.
- 7Pelckmans, M.; Vermandel, W.; Van Waes, F.; Moonen, K.; Sels, B. F. Low-temperature reductive aminolysis of carbohydrates to diamines and aminoalcohols by heterogeneous catalysis. Angew. Chem. Int. Ed 2017, 56, 14540– 14544, DOI: 10.1002/anie.2017082167https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXhs1CntLrE&md5=8737fa48cad71d03a8906ede255e3f4aLow-Temperature Reductive Aminolysis of Carbohydrates to Diamines and Aminoalcohols by Heterogeneous CatalysisPelckmans, Michiel; Vermandel, Walter; Van Waes, Frederik; Moonen, Kristof; Sels, Bert F.Angewandte Chemie, International Edition (2017), 56 (46), 14540-14544CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)Short amines, such as ethanolamines and ethylenediamines, are important compds. in today's bulk and fine chems. industry. Unfortunately, current industrial manuf. of these chems. relies on fossil resources and requires rigorous safety measures when handling explosive or toxic intermediates. Inspired by the elegant working mechanism of aldolase enzymes, a novel heterogeneously catalyzed process-reductive aminolysis-was developed for the efficient prodn. of short amines from carbohydrates at low temp. High-value bio-based amines contg. a bio-derived C2 carbon backbone were synthesized in one step with yields up to 87 C%, in the absence of a solvent and at a temp. below 405 K. A wide variety of available primary and secondary alkyl- and alkanolamines can be reacted with the carbohydrate to form the corresponding C2-diamine. The presented reductive aminolysis is therefore a promising strategy for sustainable synthesis of short, acyclic, bio-based amines.
- 8Froidevaux, V.; Negrell, C.; Caillol, S.; Pascault, J.-P.; Boutevin, B. Biobased Amines: From Synthesis to Polymers; Present and Future. Chem. Rev. 2016, 116, 14181– 14224, DOI: 10.1021/acs.chemrev.6b004868https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XhslOmu7fM&md5=970ff058a08b3941d6d727e2ffec07bfBiobased Amines: From Synthesis to Polymers; Present and FutureFroidevaux, Vincent; Negrell, Claire; Caillol, Sylvain; Pascault, Jean-Pierre; Boutevin, BernardChemical Reviews (Washington, DC, United States) (2016), 116 (22), 14181-14224CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)Amines are key-intermediates in chem. industry due to their nucleophilic characteristic which confers a high reactivity to them. Thus, they are key-monomers for the synthesis of polyamides, polyureas, polyepoxydes···which are all of growing interest in automotive, aerospace, building or health applications. Despite a growing interest for biobased monomers and polymers, and particularly polyamides, it should be noticed that very few natural amines are available. Actually, there is only chitosan and poly(lysine). In this review, we present both fundamental and applied research on the synthesis of biobased primary and secondary amines with current available biobased resources. Their use is described as building block for material chem. Hence, we first recall some background on the synthesis of amines, including the reactivity of amines. Second we focus on the synthesis of biobased amines from all sorts of biomass, from carbohydrate, from terpenes, or from oleochem. sources. Third, because they need optimization and technol. developments, we discuss some examples of their use for the creation of biobased polymers. We conclude on the future of the synthesis of biobased amines and their use in different applications.
- 9Brun, N.; Hesemann, P.; Esposito, D. Expanding the biomass derived chemical space. Chem. Sci. 2017, 8, 4724– 4738, DOI: 10.1039/c7sc00936d9https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXmsFemtLw%253D&md5=8d9a951ab73aa9488961bc3e312330fbExpanding the biomass derived chemical spaceBrun, Nicolas; Hesemann, Peter; Esposito, DavideChemical Science (2017), 8 (7), 4724-4738CODEN: CSHCCN; ISSN:2041-6520. (Royal Society of Chemistry)A review. Biorefinery aims at the conversion of biomass and renewable feedstocks into fuels and platform chems., in analogy to conventional oil refinery. In the past years, the scientific community has defined a no. of primary building blocks that can be obtained by direct biomass decompn. However, the large potential of this "renewable chem. Space" to contribute to the generation of value added bio-active compds. and materials still remains unexplored. In general, biomass derived building blocks feature a diverse range of chem. functionalities. In order to be integrated into value-added compds., they require addnl. functionalization and/or covalent modification thereby generating secondary building blocks. The latter can be thus regarded as functional components of bio-active mols. or materials and represent an expansion of the renewable chem. space. This perspective highlights the most recent developments and opportunities for the synthesis of secondary biomass derived building blocks and their application to the prepn. of value added products.
- 10Liang, G.; Wang, A.; Li, L.; Xu, G.; Yan, N.; Zhang, T. Production of primary amines by reductive amination of biomass-derived aldehydes/ketones. Angew. Chem., Int. Ed. 2017, 56, 3050– 3054, DOI: 10.1002/anie.20161096410https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXitVKntbo%253D&md5=26b501e0c8dedcc757e9655af485d6e2Production of Primary Amines by Reductive Amination of Biomass-Derived Aldehydes/KetonesLiang, Guanfeng; Wang, Aiqin; Li, Lin; Xu, Gang; Yan, Ning; Zhang, TaoAngewandte Chemie, International Edition (2017), 56 (11), 3050-3054CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)Transformation of biomass into valuable nitrogen-contg. compds. is highly desired, yet limited success has been achieved. Here we report an efficient catalyst system, partially reduced Ru/ZrO2, which could catalyze the reductive amination of a variety of biomass-derived aldehydes/ketones in aq. ammonia. With this approach, a spectrum of renewable primary amines was produced in good to excellent yields. Moreover, we have demonstrated a two-step approach for prodn. of ethanolamine, a large-market nitrogen-contg. chem., from lignocellulose in an overall yield of 10 %. Extensive characterizations showed that Ru/ZrO2-contg. multivalence Ru assocn. species worked as a bifunctional catalyst, with RuO2 as acidic promoter to facilitate the activation of carbonyl groups and Ru as active sites for the subsequent imine hydrogenation.
- 11Flèche, G.; Huchette, M. Preparation, properties and chemistry. Starch/Stärke 1986, 38, 26– 30, DOI: 10.1002/star.1986038010711https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL28XktVKksL8%253D&md5=8096fd2618fb0f588ac5c2f2fb79b799Isosorbide. Preparation, properties and chemistryFleche, G.; Huchette, M.Starch/Staerke (1986), 38 (1), 26-30CODEN: STARDD; ISSN:0038-9056.Isosorbide (I) is obtained by dehydration of sorbitol and therefore can be considered as a valuable product from biomass. The acid-catalyzed reaction gives rise to different anhydro-compds., but also to polymer-like products. Kinetics of sorbitol decrease, followed with the help of HPLC, shows, remarkedly, the different reactions taking place during the dehydration. Physicochem. properties of isosorbide are also discussed: melting temp., sp. gr., soly. Emphasis is put on stereochem. aspect, pointing out the endo-exo position of hydroxyl groups vs. those of other isomers: exo-exo for isoiodide and endo-endo for isomannide. Some other properties such as: hygroscopicity ant thermal stability are also discussed. The chem. reactivity of the mols. is described and some reactions analyzed, proving the interest of the cyclic conformation. Finally, known applications are presented.
- 12Wiggins, L. F. 2. The anhydrides of polyhydric alcohols. Part I. The constitution of isomannide. J. Chem. Soc. 1945, 4– 7, DOI: 10.1039/jr945000000412https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaH2MXhslGlsg%253D%253D&md5=c6aa485bc95f2a1019a360b1527918daAnhydrides of polyhydric alcohols. I. Constitution of isomannideWiggins, L. F.Journal of the Chemical Society (1945), (), 4-7CODEN: JCSOA9; ISSN:0368-1769.Mannitol (I) (20 g.) in 70 cc. dichloroglycerol, ClCH2CH(OH)CH2Cl heated at 140-50° while HCl is passed through the suspension for 3.5 hrs., gives 5.4 g. of 1,4,3,6-dianhydromannitol (isomannide) (II), m. 87-8°, [α]D 91° (H2O, c 2.28). I (500 g.) in 3 l. fuming HCl, boiled 72 hrs., gives 140 g. of II. 1,6-Dichloromannitol, ClCH2[CH(OH)]4CH2Cl (III) (C.A. 38, 2628.6) (2 g.), heated in vacuo at 180°, gives 0.95 g. of II. II (4 g.) in 5.6 g. C5H5N, treated with SOCl2 at 0° and heated 0.5 hr. at 100°, gives 70% of the 2,5-di-Cl compd. (IV), m. 67°, [α]D 93.5° (CHCl3, c 2.054); this is unchanged on distn. with fused KOH. II (1 g.), treated 4 times with MeI and Ag2O at 45°, gives 0.9 g. of the 2,5-di-Me ether (IVA), m. 75-6°, [α]D18 175° (CHCl3, c 2.287). II (1 g.) in 20 cc. fuming HCl, heated 24 hrs. in a sealed tube, gives 0.25 g. of III; thus positions 1 and 6 are involved in the anhydro rings. II (1 g.), heated at 120° for 29 hrs. with 50 cc. MeOH satd. with NH3 at 0°, gives 0.8 g. unchanged II; it is also unaffected with 10% MeOH-MeONa under the same conditions. II is not oxidized by HNO3 (d. 1.14) on heating at 100° for 4 hrs. Pb(OAc)4 in AcOH is without action on II in 48 hrs., indicating that the 2 OH groups are not on adjacent C atoms. IV (7.2 g.) in 200 cc. fuming HCl, heated in a sealed tube for 72 hrs. at 110°, gives 1.6 g. of unchanged IV and 3.4 g. of 1,2,5,6-tetrachloromannitol (V), m. 69-70°, [α]D 28.3° (CHCl3, c 3.107); 3,4-di-Bz deriv., m. 109-10°, [α]D17 -95.4° (CHCl3, c 1.048). V and PCl5 at 130° for 1 hr. give hexachloromannitol; thus V possesses the configuration of I. With MeONa at room temp. for 4 hrs. 1 g. of V gives 0.5 g. of IV. V (0.3 g.), shaken overnight with 25 cc. Me2CO contg. 0.1 cc. H2SO4, gives 0.33 g. of the 3,4-acetone deriv. (VI), b0.05 115° (bath temp.), nD16 1.4954, [α]D 56.8° (CHCl3, c 2.405); VI results also in 0.4-g. yield from 1 g. of 1,6-dichloro-3,4-acetonemannitol (VII)(Micheel, C.A. 26, 4304) and SOCl2 in C5H5N (1 hr. at 100°). VII (5 g.), treated 5 times with MeI and Ag2O at 45° for 9 hrs., gives 1.6 g. of the 2,5-di-Me ether (VIII), m. 56°, [α]D 11.5° (CHCl3, c 2.253); 0.503 g. of VIII in 40 cc. of 75% EtOH contg. 5% H2SO4, kept at room temp. for 550 hrs. (change in [α]D from 9.5° to -35°), gives 0.21 g. of 1,6-dichloro-2,5-dimethylmannitol, m. 131°; with MeONa in MeOH 0.17 g. gives 0.09 g. of IVA; this shows that the HO groups of II are located at C2 and C5. Mannitan (IX) distils at 10 mm. practically without decompn.; distn. of 2 g. of IX with 2 drops of concd. H2SO4 at 10 mm. gives 1.1 g. of II. IX (5 g.) in 50 cc. C5H5N, treated at 0° with 6 g. p-MeC6H4SO2Cl and kept at room temp. for 96 hrs. (with addn. of 25 cc. Ac2O after 48 hrs.), gives 10.5 g. of 1-tosyl-2,4,5-triacetylmannitan (X), a liquid. Reaction of X with MeONa gives II. That the tosyl group is on a primary alc. group is shown by treatment of X with NaI in Me2CO to give 80% of p-MeC6H4SO3Na. The above facts prove that the rings are hydrofuranol in type.
- 13Rose, M.; Palkovits, R. Isosorbide as a renewable platform chemical for versatile applications – quo vadis?. ChemSusChem 2012, 5, 167– 176, DOI: 10.1002/cssc.20110058013https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XitVamsQ%253D%253D&md5=e5a646b04db3b90495d4256201d3be04Isosorbide as a Renewable Platform chemical for Versatile Applications-Quo Vadis?Rose, Marcus; Palkovits, ReginaChemSusChem (2012), 5 (1), 167-176CODEN: CHEMIZ; ISSN:1864-5631. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. Isosorbide is a platform chem. of considerable importance for the future replacement of fossil resource-based products. Applications as monomers and building blocks for new polymers and functional materials, are conceivable. This minireview deals with all aspects of isosorbide chem., which includes its prodn., special properties, and chem. transformations for its utilization in biogenic polymers and other applications of interest.
- 14Fenouillot, F.; Rousseau, A.; Colomines, G.; Saint-Loup, R.; Pascault, J. P. Polymers from renewable 1,4:3,6-dianhydrohexitols (isosorbide, isomannide and isoidide): a review. Progr. Polym. Sci. 2010, 35, 578– 622, DOI: 10.1016/j.progpolymsci.2009.10.00114https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXjvFKitr4%253D&md5=0ca3e0de11d9077e508554de0e4ac596Polymers from renewable 1,4:3,6-dianhydrohexitols (isosorbide, isomannide and isoidide): A reviewFenouillot, F.; Rousseau, A.; Colomines, G.; Saint-Loup, R.; Pascault, J.-P.Progress in Polymer Science (2010), 35 (5), 578-622CODEN: PRPSB8; ISSN:0079-6700. (Elsevier Ltd.)A review. The use of the 1,4:3,6-dianhydrohexitols isosorbide, isomannide, and isoidide in polymers is reviewed. 1,4:3,6-Dianhydrohexitols are derived from renewable resources from cereal-based polysaccharides. 1,4:3,6-Dianhydrohexitols are derived from renewable resources from cereal-based polysaccharides. In the field of polymeric materials, these diols are essentially employed to synthesize or modify polycondensates. Their attractive features as monomers are linked to their rigidity, chirality, non-toxicity, and the fact that they are not derived from petroleum. First, the synthesis of high glass transition temp. polymers with good thermomech. resistance is possible. Second, the chiral nature of 1,4:3,6-dianhydrohexitols may lead to specific optical properties. Finally, biodegradable polymers can be obtained. The prodn. of isosorbide on an industrial scale with a purity satisfying the requirements for polymer synthesis suggests that isosorbide will soon emerge in industrial polymer applications. However, a deciding factor will be the redn. of polymn. time of these low-reactivity monomers to values compatible with economically viable prodn. processes to give polyesters, polyamides, poly(amide esters), poly(ester imides), polycarbonates, polyurethanes, and polyethers.
- 15Wu, J.; Jasinska-Walc, L.; Dudenko, D.; Rozanski, A.; Hansen, M. R.; van Es, D.; Koning, C. E. An investigation of polyamides based on isoidide-2,5-dimethyleneamine as a green rigid building block with enhanced reactivity. Macromolecules 2012, 45, 9333– 9346, DOI: 10.1021/ma302126b15https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38Xhs12mtLrE&md5=7322665bbe1faba162bad35a629950adAn Investigation of Polyamides Based on Isoidide-2,5-dimethyleneamine as a Green Rigid Building Block with Enhanced ReactivityWu, Jing; Jasinska-Walc, Lidia; Dudenko, Dmytro; Rozanski, Artur; Hansen, Michael Ryan; van Es, Daan; Koning, Cor E.Macromolecules (Washington, DC, United States) (2012), 45 (23), 9333-9346CODEN: MAMOBX; ISSN:0024-9297. (American Chemical Society)Novel, semicryst. polyamides and copolyamides were synthesized from a new carbohydrate-based diamine, namely isoidide-2,5-dimethyleneamine (IIDMA). In combination with 1,6-hexamethylene diamine (1,6-HDA) as well as the biobased sebacic acid (SA) or brassylic acid (BrA), the desired copolyamides were obtained via melt polymn. of the nylon salts followed by a solid-state polycondensation (SSPC) process. Depending on the chem. compns., the no. av. mol. wts. (Mn) of the polyamides were in the range of 4000-49000 g/mol. With increasing IIDMA content in the synthesized copolyamides, their corresponding glass transition temps. (Tg) increased from 50 °C to approx. 60-67 °C while the melting temps. (Tm) decreased from 220 to 160 °C. The chem. structures of the polyamides were analyzed by NMR and FT-IR spectroscopy. Both differential scanning calorimetry (DSC) and wide-angle X-ray diffraction (WAXD) analyses revealed the semicryst. character of these novel copolyamides. Variable-temp. (VT) 13C{1H} cross-polarization/magic-angle spinning (CP/MAS) NMR and FT-IR techniques were employed to study the crystal structures as well as the distribution of IIDMA moieties over the cryst. and amorphous phases of the copolyamides. The performed ab initio calcns. reveal that the stability of the IIDMA moieties is due to a pronounced "boat" conformation of the bicyclic rings. The incorporation of methylene segments in between the isohexide group and the amide groups enables the hydrogen bonds formation and organization of the polymer chain fragments. Given the sufficiently high Tm values (∼200 °C) of the copolyamides contg. less than 50% of IIDMA, these biobased semicryst. copolyamides can be useful for engineering plastic applications.
- 16Montgomery, R.; Wiggins, L. F. 77. The anhydrides of polyhydric alcohols. Part IV. The constitution of dianhydro sorbitol. J. Chem. Soc. 1946, 390– 393, DOI: 10.1039/jr946000039016https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaH28XisFKmug%253D%253D&md5=f1e77560ace47ba4790e450f6302f844Anhydrides of polyhydric alcohols. IV. Constitution of dianhydrosorbitolMontgomery, R.; Wiggins, L. F.Journal of the Chemical Society (1946), (), 390-3CODEN: JCSOA9; ISSN:0368-1769.cf. Hockett, et al., C.A. 40, 4677.3. Sorbitol (500 g.) and concd. HCl, refluxed 24 hrs., give 265 g. (66%) of 1,4,3,6-dianhydrosorbitol (I), b10 160-5°, m. 61-3°, [α]D 43.9° (H2O, c 0.8). I does not react with Pb(OAc)4 in AcOH or with BzH in the presence of ZnCl2. I (100 g.) in 200 cc. H2O at 40°, treated with 258 g. Me2SO4 and 360 cc. 30% NaOH in 10 portions during 2 hrs., gives 92 g. of the 2,5-di-Me deriv. (II), b0.1 93-5° (bath), nD19 1.4622, [α]D21 92.9° (CHCl3, c 1.51). I does not react with 5% MeONa in MeOH (20 hrs. at 120°) or with MeOH-NH3 (satd. at 0°) at 120° for 30 hrs. I (5 g.) and 50 cc. fuming HCl, heated at 100-10° for 24 hrs., give 5 g. of a dark brown sirup which, with (HCHO)3 and concd. H2SO4, yields 1,6-dichloro-2,4,3,5-dimethylenesorbitol, m. 116° (C.A. 38, 2628.6). 1,2-Acetoneglucose (40 g.) yields 30 g. of the 5,6-ditosyl deriv. (III) and 18.2 g. of the 3,5,6-tritosyl deriv., m. 129-30°, [α]D17 -3.4° (CHCl3, c 4.14), which results also on further tosylation of III (cf., however, Ohle, et al., C.A. 23, 103, who describe a 3,5,6-deriv. m. 95-6°, [α]D -5.2°). Reduction of 3,6-anhydro-glucose over Raney Ni at 110-20°/100 atm. gives 3,6-anhydrosorbitol (IV); distn. of 0.2 g. with a trace of H2SO4 gives 25 mg. of I. IV (1 g.) and 1.28 g. p-MeC6H4-SO2Cl in 10 cc. C5H5N, mixed at 0° and kept at room temp. for 60 hrs., Ac2O being added at 0° after the 1st 16 hrs., give 1.63 g. of 1-tosyl-2,4,5-triacetyl-3,6-anhydrosorbitol (V); 1.2 g. of V and 0.25 g. Na in 20 cc. MeOH and 10 cc. CHCl3, on standing overnight, give 73% of I. V (0.4 g.) and 0.29 g. NaI in 15 cc. Me2CO, heated at 110°, give 80% of p-MeC6H4SO3Na; the other product could not be crystd. Catalytic reduction of 1 g. of 2,5-dimethyl-3,6-anhydroglucose in H2O over 1 g. Raney Ni at 110-20°/100 atm. for 6 hrs., gives 1 g. of 2,5-dimethyl-3,6-anhydrosorbitol, m. 70-1°, [α]D22 -15.6° (CHCl3, c 0.173); treated as above with p-MeC6H4SO2Cl and Ac2O in C5H5N, 0.3 g. yields 0.23 g. of 1-tosyl-4-acetyl-2,5-dimethyl-3,6-anhydrosorbitol which, with MeONa in MeOH-CHCl3, gives 0.07 g. of II.
- 17Fletcher, H. G., Jr; Goepp, R. M., Jr Hexitol Anhydrides.1 1,4,3,6-Dianhydro-L-iditol and the Structures of Isomannide and Isosorbide. J. Am. Chem. Soc. 1946, 68, 939– 941, DOI: 10.1021/ja01210a00717https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaH28Xitl2jsQ%253D%253D&md5=b78bfdb5c19e6d0887e57a9ffd49914d1,4,3,6-Dianhydro-L-iditol and the structure of isomannide and isosorbideFletcher, Hewitt G., Jr.; Goepp, R. Max, Jr.Journal of the American Chemical Society (1946), 68 (), 939-41CODEN: JACSAT; ISSN:0002-7863.1,4,3,6-Dianhydrosorbitol (isosorbide) (I) (50 g.) and 10 g. Raney Ni in abs. EtOH, freed of EtOH in vacuo, heated 10 min. at 135-40°/2 mm., and then subjected to a pressure of 3750 lb./sq. in. of H at 190-200° for 2 hrs., gave a sirup which could not be crystd. Distn. of 46.1 g. at 140° (bath)/2 mm. gave 27.2 g. of distillate, [α]D25 44.2° in AcOH, and 18.1 g. of a residue (II), [α]D25 27.4° in AcOH. II (18.1 g.) in 50 ml. C5H5N at 0°, treated with 32 ml. BzCl, allowed to stand overnight at room temp., poured into 726 ml. ice water, and crystd. from EtOH, gave 31 g. of a product, m. 81.8-90°; the CHCl3, soln., shaken with aq. NaHCO3 to remove the BzOH and the product crystd. from EtOH and BuOH, gave 16.1 g. of the 2,5-di-Bz deriv. (III), m. 111-11.3° (m. ps. cor.), [α]D23.4 140.3° (CHCl3, c 2.03), [α]D28.2 110.5° (C5H5N, c 2.07), of 1,4,3,6-dianhydro-L-iditol (L-isoidide) (IV), m. 63.8-4.4°, [α]D24.5 20.8° (CHCl3, c 2.02), [α]D28.2 33.3° (C5H5N, c 2.24°). By the same method, 50 g. of isomannide (V) gives 0.92 g. of III. L-Iditol (0.97 g.) and 2 drops concd. H2SO4, heated at 140-5° (bath)/4 cm. for 1.25 hrs. and the product treated with BzCl in C5H5N, gave 31.4% of III. III (10 g.) in 40 ml. CHCl3 at 0°, treated with a chilled soln. of 0.1 g. Na in 40 ml. MeOH and kept at 0° for 23 hrs., gave 90.4% of IV, which is sparingly sol. in CHCl3. The prepn. of IV is best explained by the assumption that Raney Ni exerts a dehydrogenating action on the secondary alcs., converting 1 or both of the free HO groups to sym. CO groups, which are subsequently reduced with the formation of a mixt. of diastereoisomers. The isolation of IV from I and V proves the structure of IV and confirms those of I and V.
- 18Fletcher, H. G., Jr; Goepp, R. M., Jr 1,4;3,6-Hexitol dianhydride L-isoidide. J. Am. Chem. Soc. 1945, 67, 1042– 1043, DOI: 10.1021/ja01222a51318https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaH2MXitVKktw%253D%253D&md5=a4fed19fe54813233d8dcd51adc301491,4,3,6-Hexitol dianhydride. L-IsoidideFletcher, Hewitt G., Jr.; Goepp, R. Max, Jr.Journal of the American Chemical Society (1945), 67 (), 1042-3CODEN: JACSAT; ISSN:0002-7863.Reduction of D-isomannide or D-isosorbide at 200° over Raney Ni at 250 atms. gives a mixt. of hexitol dianhydrides from which, by benzoylation and fractional crystn., there is sepd. 1,4,3,6-dianhydro-L-iditol (L-isoidide) (I), m. 63.7-4.5°, [α]D24.5 20.8° (H2O, c 2.02), [α]D28.2 33.27° (C5H5N, c 2.24); dibenzoate, m. 111-11.3°, [α]D25.2 141.9° (CHCl3, c 2.15), [α]D28.2 110.5° (C5H5N, c 2.07). Similar data are given for the D-sorbitol and D-mannitol derivs. I also results by the direct acid-catalyzed anhydrization of L-iditol.
- 19Cope, C.; Shen, T. Y. The stereochemistry of 1,4: 3,6-dianhydrohexitol derivatives. J. Am. Chem. Soc. 1956, 78, 3177– 3182, DOI: 10.1021/ja01594a05519https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaG28Xnslyjsg%253D%253D&md5=1703546f35eedae5c3cca190ea364ca6Stereochemistry of 1,4:3,6-dianhydrohexitol derivativesCope, Arthur C.; Shen, T. Y.Journal of the American Chemical Society (1956), 78 (), 3177-82CODEN: JACSAT; ISSN:0002-7863.Ditosylate (I) (35 g.) of D-isomannide (II) and 40 g. Et4NOAc.H2O (III) refluxed 70 hrs. on the steam bath in 400 cc. Me2CO, concd. to 100 cc., dild. with 400 cc. H2O, and extd. with CHCl3, the extd. evapd., the residue distd., and the distillate (b0.7 120-50° bath) cooled gave 14.1 g. diacetate (IV) of 1,4:3,6-dianhydro-L-iditol (V), b0.5 100-10°, m. 57-7.6°, [α]D25 89.6° (c 1.5, CHCl3); all m.ps. are cor. II (4.0 g.) treated 2 days at 0-5° with 12 cc. Ac2O and 15 cc. pyridine, dild. with H2O, and extd. with CHCl3 gave 5.63 g. diacetate (VI) of 1,4:3,6-dianhydro-D-mannitol (VII), b0.5 118°, nD25 1.4680, [α]D25 194.5° (c 6.7, CHCl3). IV (12.5 g.) in 90 cc. abs. MeOH treated at 0° with 10 cc. N Ba(OMe)2 in MeOH, kept 18 hrs. at 5°, treated with powd. Dry Ice, and centrifuged, and the supernatant liquid worked up yielded 7.1 g. V, m. 43-3.5°, [α]D25 18.7° (c 2, H2O). IV (350 mg.) in 5 cc. 20% KOH shaken 20 min. with 1 cc. BzCl yielded 386 mg. dibenzoate of V, flakes, m. 110.6-11.4° (from aq. EtOH), [α]D26 134.2° (c 3, CHCl3). V (1.46 g.) treated 18 hrs. at 0-5° with 4.5 g. p-MeC6H4SO2Cl in 15 cc. pyridine and dild. with ice water yielded 3.96 g. di-p-toluene-sulfonate (VIII) of V, m. 105.5-106° (from MeOH), [α]D25 33.2° (c 2.5, CHCl3). Ditosylate (IX) (9.0 g.) of isosorbide refluxed 72 hrs. with 5.0 g. III in Me2CO, concd., dild. with H2O, and extd. with CHCl3 yielded 4.54 g. 5-O-acetyl-2-O-p-toluenesulfonyl-1,4:3,6-dianhydro-L-iditol (X), m. 95.5-6.3° (from MeOH), [α]D25 50.5° (c 4.7, CHCl3). X (1.0 g.) in 10 cc. abs. MeOH treated at 0° with 10 cc. 0.1N Ba(OMe)2 in MeOH, neutralized after 18 hrs. with powd. Dry Ice, centrifuged, and evapd. in vacuo, the oily residue treated 18 hrs. with 1.0 g. p-MeC6H4SO2Cl in 10 cc. pyridine at 5°, and the mixt. poured into ice water and filtered yielded 0.67 g. VIII, m. 105-6°. VIII (3.8 g.) and 1.75 g. III in 50 cc. Me2CO refluxed 72 hrs., concd. to 20 cc. and dild. with H2O gave 2.19 g. unchanged VIII, m. 104.5-5.3°. I (22.6 g.), 20 g. K phthalimide, and 300 cc. HCONMe2 heated 40 hrs. at 110° and dild. with 1.5 l. H2O, and the pptd. solid extd. with 300 cc. boiling EtOH left 6.9 g. 2,5-diphthalimido-2,5-dideoxy deriv. (XI) of V, prisms, m. 243.4-3.6° (from EtOAc-EtOH), [α]D25 168° (c 1, CHCl3). XI (8.08 g.), 1.51 g. 85% N2H4.H2O, and 200 cc. EtOH refluxed 2 days, treated with 20 cc. 4N HCl, refluxed 1 hr., and filtered, the filtrate evapd. to dryness in vacuo, the residue dissolved in 30 cc. H2O, filtered, treated with Darco, acidified to pH about 4.5 with aq. (CO2H)2, and dild. with EtOH to ppt. 2.1 g. oxalate, m. 242-3°, and the oxalate dissolved in H2O, treated with 0.7 g. NaOH in H2O, and distd. yielded 0.97 g. 2,5-di-NH2 analog (XII) of XI, b0.2 110°, m. 59-60°; picrate, m. 227.8-28.1°. I (39 g.) and 70 g. Me2NH shaken 72 hrs. at 120° in 500 cc. tetrahydrofuran in a steel bomb, the mixt. concd., treated with 100 cc. 20% aq. NaOH and extd. with Et2O, the ext. evapd., and the sirupy residue distd. yielded 7.5 g. 2,5-di-(Me2NCH2) analog (XIII) of XI, b0.2 80-100°, m. 57.5-8.5° (sublimed at 40° and 0.2 mm.), [α]D26 30.0° (c 2.1, H2O). XIII treated with excess MeI in boiling MeOH and recrystd. from aq. EtOH gave XIII.2MeI.H2O, m. 260° (decompn.), [α]D26 33.3° (c 2.2, H2O). XIII gave a dipicrate, m. 222° (decompn.) (from aq. EtOH). IX (45.2 g.) and 100 g. Me2NH in 600 cc. tetrahydrofuran heated at 120-30° with shaking in a bomb, cooled, concd., treated with 150 cc. 20% NaOH, and extd. with Et2O, and the ext. worked up gave 0.68 g. crude low boiling fraction, b0.3 41°, and 14.5 g. residue; the distillate treated with MeI in MeOH yielded 1.13 g. 5-dimethylamino-5-deoxy-1,4:3,6-dianhydro-1,2-L-iditoleen methiodide, plates, m. 202-3° (decompn.), [α]D31 33.1° (c 2.5, H2O); the nonvolatile residue in Et2O washed with H2O, dried over KOH, and concd. gave a brown sirup, which treated with picric acid yielded the picrate of 2-O-p-toluenesulfonyl-5-dimethylamino-5-deoxy-1,4:3,6-dianhydro-L-iditol (XIV), m. 179.4-80.8°. The sirup treated with MeI in MeOH gave XIV.MeI, prisms, m. 177.6-8.6° (from EtOH). IX (45.2 g.) and 100 g. Me2NH in 600 cc. tetrahydrofuran shaken 48 hrs. at 165°, concd., treated with 150 cc. 20% aq. NaOH, and extd. with Et2O, and the ext. worked up yielded 13.4 g. 2,5-bis(dimethylamino)-2,5-dideoxy-1,4:3,6-dianhydro-D-glucitol (XV), white prisms, m. 54.6-5.4° (sublimed at 80° and 0.5 mm.), [α]D25 106.3° (c 1.6, H2O). Cryst. oxalate (1.17 g.) of 2,5-di-NH2 analog (XVI) of XV in 10 cc. 98% HCO2H and 8 cc. 37% CH2O refluxed 18 hrs., treated with 10 cc. 4N HCl and evapd. to dryness in vacuo, the residue dissolved in 30 cc. N NaOH and extd. with Et2O, the ext. worked up, and the residue sublimed at 80° and 0.5 mm. gave 0.67 g. XV, m. 54-4.5°, which gave XV.2MeI.H2O, m. 292-5° (decompn.). XVI (3.0 g.) and 25 g. MeI in 100 cc. abs. MeOH refluxed 48 hrs. with stirring with 4 g. Na2CO3, acidified with 48% HI, and evapd. to dryness in vacuo yielded 9.25 g. XVI.2MeI, m. 289-94° (decompn.). XVI.2MeI in 100 cc. dry 1-methylmorpholine treated slowly with 2 g. LiAlH4 in 20 cc. 1-methylmorpholine with stirring and cooling, the mixt. heated 48 hrs. at 90-5°, treated with cooling with EtOAc, and centrifuged, the ppt. dissolved in 10% NaOH and extd. with Et2O, and the combined ext. and 1-methylmorpholine soln. fractionated gave 2.73 g. XV, m. 51.5-3.5°. XV (1.0 g.) warmed 15 min. on the steam bath with 2.0 g. p-MeC6H4SO2Cl in 20 cc. 10% aq. NaOH gave only 135 mg. pure XV, m. 54.3-5.3°. XII (0.72 g.) in 10 cc. 98% HCO2H and 6 cc. 37% CH2O refluxed 18 hrs. gave 0.72 g. XIII, m. 52-4°. XII (0.320 g.) and 5 g. MeI in 20 cc. abs. MeOH refluxed 65 hrs. with 0.85 g. NaHCO3 and worked up in the usual manner yielded 0.667 g. XIII.2MeI, prisms, m. 225-8° (decompn.) (from aq. EtOH). XII (0.72 g.) and 3 cc. concd. HCl in 30 cc. glacial AcOH treated slowly with cooling and stirring with 2.0 g. BuONO, the soln. kept 0.5 hr. at 0° and 18 hrs. at room temp., concd. in vacuo, dild. with H2O, and extd. with CHCl3, the ext. worked up, the residual sirup distd., the resulting oil chromatographed on 10 g. Al2O3 with 5:1 petr. ether-Et2O, and the 1st 60 cc. effluent worked up gave 103 g. XVII (R = R' = Cl), m. 67° XVI (0.72 g.) deaminated in the same manner yielded 87 mg. XVII. On the basis of these results D-isomannide has 2 endo-OH groups, isosorbide has 1 endo at C-5 and 1 exo at C-2, and L-isoidide has 2 exo-OH groups. Revised configurations are suggested for a no. of dianhydrohexitol derivs. and other compds. contg. bicyclic ring systems with 2 cisfused 5-membered rings on the basis of the steric effects observed (previous configuration and revised configuration given): 5-chloro-5-deoxy(?)-1,4:3,6-dianhydro-D-glucitol (W. G. Overend, et al., C.A. 43, 2939f), 5-chloro-5-deoxy-1,4:3,6-dianhydro-L-iditol; x-iodo-x'-O-p-toluenesulfonyl-1,4:3,6-dianhydro-D-glucitol (Hockett, et al., C.A. 40, 4677.3), 5-iodo-5-deoxy-2-O-p-toluenesulfonyl-1,4:3,6-dianhydro-L-iditol; 2,5-dichloro-2,5-dideoxy-1,4:3,6-dianhydro-D-mannitol (Wiggins, C.A. 39, 2736.9), 2,5-dichloro-2,5-dideoxy-1,4:3,6-dianhydro-L-iditol; 2-chloro-2-deoxy-5-O-methanesulfonyl-1,4:3,6-dianhydro-D-mannitol (Montgomery and Wiggins, C.A. 43, 2940a), 2-chloro-2-deoxy-5-O-methanesulfonyl-1,4:3,6-dianhydro-D-glucitol; 2-chloro-2-deoxy-5-phenylcarbamyl(?)-1,4:3,6-dianhydro-D-mannitol (Carr´e and Mauclere, C.A. 25, 4526), 2-chloro-2-deoxy-5-phenylcarbamyl-1,4:3,6-dianhydro-D-glucitol; 2,5-diiodo-2,5-dideoxy-1,4:3,6-dianhydro-D-mannitol (Hockett, et al., C.A. 40, 4677.8), 2,5-diiodo-2,5-dideoxy-1,4:3,6-dianhydro-L-iditol; 2,5-diamino-2,5-dideoxy-1,4:3,6-dianhydro-D-mannitol and derivs. (Montgomery and Wiggins, C.A. 40, 5016.4), 2,5-diamino-2,5-dideoxy-1,4:3,6-dianhydro-L-iditol and derivs.; 2,5-dithio-2,5-dideoxy-1,4:3,6-dianhydro-D-mannitol and derivs. (Bladon and Owen, C.A. 44, 6811h), 2,5-dithio-2,5-dideoxy-1,4:3,6-dianhydro-L-iditol and derivs.
- 20Brimacombe, J.; Foster, A. B.; Stacey, M.; Whiffen, D. H. Aspects of stereochemistry – I: Properties and reactions of some diols. Tetrahedron 1958, 4, 351– 360, DOI: 10.1016/0040-4020(58)80056-320https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaG1MXmslGgsg%253D%253D&md5=002e6c00694f528888914553ddabe043Aspects of stereochemistry. I. Properties and reactions of diolsBrimacombe, J. S.; Foster, A. B.; Stacey, M.; Whiffen, D. H.Tetrahedron (1958), 4 (), 351-60CODEN: TETRAB; ISSN:0040-4020.The extent of intramol. H-bonding, as detd. by infrared spectroscopy in the HO stretching region in certain vicinal diols provided evidence for the stabilities of certain conformations. With the exception of cyclohexane-trans-1,2-diol (I), the carbocyclic trans diols were prepd. by bromination of the corresponding olefins, solvolysis of the trans bromides with AcOH and AgOAc and sapon. of the trans diacetates (Winstein and Roberts, C.A. 48, 5812f). The carbocyclic cis diols were obtained by hydroxylation of the appropriate olefins with KMnO4 (Clarke and Owen, C.A. 43, 7434h). Cycloheptene (12 g.) in 40 ml. Et2O at 0° brominated in Et2O, the colored soln. washed successively with 0.01N NaOH and water, the dried (MgSO4) soln. evapd. and the residue distd. gave 25.4 g. trans-1,2-dibromocycloheptane b20-5 138-40°, n23.5D 1.5532; this added to 40 g. AgOAc in 80 ml. AcOH and 25 ml. Ac2O (previously kept 2 hrs. at 110°), the mixt. kept 15 hrs. at 110° and the filtered soln. evapd. gave 8.2 g. trans-1,2-diacetoxycycloheptane b0.5 99-101°, n23D 1.4530. The diacetate (4 g.) boiled 2 hrs. in 8 ml. alc. and 8 ml. 35% aq. NaOH, the mixt. dild. with 8 ml. water and extd. 1 day with Et2O yielded 74% cycloheptane-trans-1,2-diol (II), b0.5 100°, m. 60-2°. Similarly were prepd. indan-trans-1,2-diol (III), m. 156-7°, and cyclopentane-trans-1,2-diol (IV), b15-20 131°. I, m. 102-3°, was prepd. according to Roebuck and Adkins [Org. Syntheses, Collective Vol. III, 217(1955)]. The cis-1,2-diols of cyclohexane (V), m. 93-4°, cycloheptane (VI), m. 45-6°, cyclopentane (VII), b29 133°, and indan (VIII), m. 96-8°, were obtained in 14, 11, 11, and 10.8-14.5% yields, resp. 3,4-Di-O-acetyl-D-xylal [3 g., m. 36-8°, [α]20D -288° (c 1.0, H2O)] prepd. according to Overend, et al. (C.A. 44, 6819f) in 100 ml. 1:1 alc.-H2O hydrogenated 30 min. with 100 mg. PtO2, the filtered soln. evapd. and the product distd. gave 1.3 g. 3,4-di-O-acetyl-1,5-anhydro-2-deoxy-D-threopentitol (IX), b0.5 102°, [α]20D -38° (c 0.85, CHCl2), [M]D -77°. IX (1.4 g.) refluxed 2 hrs. in 8 ml. 6N NaOH and 8 ml. alc., the soln. filtered through Amberlite IR-120 (H+) and the filtrate evapd. yielded 60.7% 1,5-anhydro-2-deoxy-D-threopentitol (X), b0.5 75-80°, m. 69°, [α]20D -29.6° (c 2.5, H2O), [M]D -35°. Hydrogenation of 3.2 g. 3,4-di-O-acetyl-L-arabinal (b0.01 110-40°, [α]20D -236° (c 1.12, CHCl3), prepd. according to Deriaz, et al. (C.A. 44, 2453a), and worked up similarly to X gave 2.2 g. 3,4-di-O-acetyl-1,5-anhydro-2-deoxy-L-erythropentitol (XI), b0.2 86-90°, [α]20D 75° (c 1.0, H2O), [M]D 151°. XI (1.4 g.) sapond. as for IX yielded 60.7% 1,5-anhydro-2-deoxy-L-erythropentitol (XII), b0.2-0.3 120°, [α]20D 64° (c 1.5, H2O), [M]D 75°. L-Arabinal (0.8 g., m. 80-2°, [α]20D -202° (c 3.0, H2O) prepd. according to D., et al.) in 20 ml. 1:1 alc.-H2O hydrogenated 1 hr. with 50 mg. PtO2 and the filtered soln. evapd. yielded 70.4% XII. DiBu L-tartrate (10 g., n16.5D 1.4450, [α]20D 10.2°) in 50 ml. tetrahydrofuran added slowly with stirring to 6 g. LiAlH4 in 150 ml. tetrahydrofuran and 75 ml. Et2O, the mixt. refluxed 1.5 hrs. and decompd. with 200 ml. H2O, the mixt. centrifuged and the solid residue washed with water, the combined centrifugates evapd. and the residue taken up in 200 ml. 1:1 H2O-MeOH, the soln. neutralized with CO2 and the filtered soln. evapd., the glassy solid taken up in 20 ml. 1:1 H2SO4-H2O and shaken 1 hr. with 15 ml. BzH, the mixt. dild. with water and filtered, the ppt. washed (water) and crystd. (PhMe) yielded 2.2 g. 1,2,3,4-di-O-benzylidene-L-threitol (Klosterman and Smith, C.A. 48, 2585f), m. 216-17°, [α]25D 81° (c 0.5, CHCl3), hydrolyzed 30 min. in 40 ml. 1:3 N H2SO4-alc., the hydrolyzate concd. and extd. with Et2O, the ext. filtered through Amberlite IRA-400 (HO-), the filtrate evapd. and the residue recrystd. to give 0.5 g. L-threitol (XIII), m. 88-9°, [α]20D -4° (c 8.0, H2O). XIII (2.17 g.) in 2.2 g. H2O and 2.2 g. H2SO4 heated 24 hrs. at 120° in a sealed tube, the hydrolyzate dild. with water, the soln. filtered through Amberlite IRA-400 (HO-) and the filtrate concd. gave 1.1 g. 1,4-anhydro-L-threitol (XIV), b15-17 120°, m. 63-4°, [α]20D -4° (c 7.2, H2O). Erythritol (3 g.) in 3 g. H2O and 3 g. H2SO4 heated 2 days at 120° and in a sealed tube and the hydrolyzate worked up as for XIII gave 1,4-anhydroerythritol (XV), b2-3 144°, n20D 1.4767. The zone electrophoretic (ionophoretic) mobility of the diols were detd. using the app. and technique of Foster, et al. (C.A. 50, 4584g), with a borate buffer (pH 10). The mobility, MG is defined as that mobility relative to D-glucose under standard conditions. The diols (0.43 millimole) in 5 ml. H2O were treated with 15 ml. 0.05M NaIO4 and the vol. rapidly adjusted to 100 ml. at 0°, aliquots withdrawn periodically titrated for unchanged NaIO4 by the standard arsenite method and where possible the times of half oxidation (t0.5) were recorded. The diols (26 mg.) in AcOH at 20° were treated with 49 ml. 0.1315N Pb(OAc)4 in AcOH at 20°, the vol. adjusted to 50 ml. and the unconsumed oxidant detd. on 5-ml. aliquots according to Hockett and McClenahan (C.A. 33, 68034). Methylation of 1,3-O-methylidene glyceritol (XVI), b11 82°, n20D 1.4533, according to Hibbert and Carter (C.A. 23, 98) gave 5-O-methyl-1,3-dioxan-5-ol, b. 147°, n25D 1.4230, converted by acid hydrolysis to 2-O-methyl-1,3-O-methylidene glyceritol, b13 120°, n23D 1.4476, completely resistant to attack by Pb(OAc)4 under the above conditions. Infrared spectra were measured in 2-cm. layers in CCl4 according to Spedding and Whiffen (C.A. 51, 15275b) with concns. of less than 0.005M diol. The results are tabulated (compd., ν for free and bonded OH in cm.-1, arithmetical difference between frequencies, arithmetical difference between standard secondary (3629 cm.-1) and bonded OH group frequencies, t0.5 for NaIO4 oxidation, t0.5 for Pb(OAc)4 oxidation and MG given): I, 3633, 3602, 31, 27, rapid, 1.9 hrs., 0.00; X, 3633, 3608, 25, 21, 20 min., 5 hrs., 0.00; V, 3632, 3592, 40, 37, rapid, 5 min., 0.07; XII, 3633, 3583, 50, 46, rapid, 9 min., 0.23; IV, 3624, -, -, -, 5 min., 10 min., 0.00; XIV, 3624, -, -, -, -, 11 hrs., 0.00; VII, 3624, 3579, 45, 50, rapid, rapid, 0.69; XV, 3624, 3585, 39, 44, -, rapid, 0.88; 1,4,3,6-dianhydro-D-glucitol, 3624, 3540, 84, 89, -, -, -; 1,4,3,6-dianhydro-D-glucitol, 3624, 3540, 84, 89, -, -, -; 1,4,3,6-dianhydro-D-glucitol, 3624, 3540, 84, 89, -, -, -; 1,4,3,6-dianhydro-D-mannitol, -, 3540, -, 89, -, -, -; 1,4,3,6-dianhydro-L-iditol, 3624, -, -, -, -, -, -; III, 3624, -, -, -, 60 min., -, 0.00; VII, 3620, 3579, 41, 50, rapid, -, 0.72; II, 3626, 3589, 37, 40, rapid, -, 0.53; VI, 3632, 3588, 44, 41, rapid, -, 0.69; XVI, 3625, 3593, 42, 36, -, -, -. In certain compds. the stabilities of different conformations can be affected by H-bonding from a substituent HO group to a ring O. Using the parameters evaluated by Whiffen (C.A. 51, 4261i) [M]D in aq. soln. were calcd. for the possible chair conformations of certain pyran derivs. (diol, conformation, [M]D calcd., and [M]D observed given): X, HO groups axial, -43°, -35°, HO groups equatorial, -45°, -; XII, C-3 HO group axial, -45°, -, C-4 HO group axial, 88°, 75°. The rate of reaction of the vicinal diols of these cyclic systems with glycol splitting reagents, and their zone electrophoretic mobility in an alk. borate buffer is influenced by the presence of a ring O atom.
- 21Montgomery, R.; Wiggins, L. F. 78. The anhydrides of polyhydric alcohols. Part V. 2 : 5-Diamino 1 : 4–3 : 6-dianhydro mannitol and sorbitol and their sulphanilamide derivatives. J. Chem. Soc. 1946, 0, 393– 396, DOI: 10.1039/jr946000039321https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaH28XisFKmuw%253D%253D&md5=ff5e9a4f4c99a95b6626adda67069ca2Anhydrides of polyhydric alcohols. V. 2,5-Diamino-1,4,3,6-dianhydromannitol and -sorbitol and their sulfanilamide derivativesMontgomery, R.; Wiggins, L. F.Journal of the Chemical Society (1946), (), 393-6CODEN: JCSOA9; ISSN:0368-1769.1,4,3,6-Dianhydromannitol (59 g.) in 300 cc. dry C5H5N, treated at 0° with 155 g. p-MeC6H4SO2Cl in portions and allowed to warm to and stand at room temp. for 24 h., gives 162 g. of the 2,5-ditosyl deriv. (I), m. 93-4°, [α]D20 92.2° (CHCl3, c 2.582). I (50 g.) in 1500 cc. MeOH-NH3 (satd. at 0°), heated at 170-80° for 30 h., the residue heated with 42 g. Ba(OH)2 in 400 cc. H2O for 1 h., the residue from this operation freed from H2O by repeated distn. with C6H6, and extd. 6 times with 250-cc. portions of CHCl3, gives 12.8 g. of 2,5-diamino-1,4,3,6-dianhydromannitol (II), a sirup, b0.01 150° (bath), m. 59-62°, [α]D20 33.6° (CHCl3, c 2.322), which could be kept in cryst. condition only under N in sealed tubes; oxalate, m. 246-7° (decompn.); adipate, m. 189°; picrate, m. 227-8° (decompn.); sulfate, decomps. above 310°; dimethylenemucate, m. 246-7° (decompn.). II did not inhibit the growth of Staphylococcus aureus in vitro. II yields a disalicylidene deriv., yellow, m. 188-9°. The bis(p-acetamidophenylsulfonyl) deriv. (III), m. 278-9°, was prepd. from II and p-AcNHC6H4SO2Cl in dil. Me2CO contg. NaOH or NaHCO3 on stirring at room temp. for 45 min. or in C5H5N at room temp. for 2 days; it is optically inactive. Hydrolysis of 10 g. of III in 100 cc. Me2CO and 200 cc. 2 N HCl (refluxing 6 h.), or by heating 1 g. with 10 cc. 10% NaOH at 100° for 2 h., gives 2,5-disulfanilamido-1,4,3,6-dianhydromannitol (IV), m. 227-8°; it is optically inactive and does not form salts in aq. soln.; dry HCl in Me2CO-C6H6 (1:1) yields a di-HCl salt which effervesces between 180° and 215°, and yields IV with cold H2O. Reacetylation of IV (AcOH-Ac2O) gives III. II (3 g.) and p-O2NC6H4SO2Cl in C5H5N yield 9.3 g. of the bis(p-nitrophenylsulfonyl) deriv. (V), m. 213-14°, [α]D20 7.5° (Me2CO, c 2.65); redn. with Sn and HCl, or over Raney Ni, gives IV. Dianhydrosorbitol (58 g.) yields 190 g. of the 2,5-ditosyl deriv. (V), m. 101-2°, [α]D23 57.8° (CHCl3, c 4.945); treated as in the case of I, 50 g. of V yields 8.5 g. (54%) of 2,5-diamino-1,4,3,6-dianhydrosorbitol (VI), b0.01 105-10° (bath), nD18 1.5165, [α]D 43.6° (H2O, c 1.538); oxalate, m. 253-4° (decompn.); picrate, m. 200° (decompn.); HCl salt, does not m. 320°; sulfate, does not m. 330°; dimethylenemucate, m. 235-6° (decompn.); dimethylenesaccharate, m. 220-1° (decompn.). The salts did not inhibit the growth of S. aureus in vitro. The disalicylidene deriv. of VI m. 186-7°; bis(p-acetamidophenylsulfonyl) deriv., m. 263-4°, [α]D 51.4° (Me2CO-H2O (1:1), c 0.6); 2,5-disulfanilamido-1,4,3,6-dianhydrosorbitol (VII), m. 239-40°, [α]D 49.2° (Me2CO, c 0.406); the di-HCl salt, prepd. under anhyd. conditions, effervesces about 130° and is completely hydrolyzed by cold H2O. Bis(p-nitrophenylsulfonyl) deriv., m. 216-17°, [α]D20 56.8° (Me2CO, c 2.138); catalytic redn. over Raney Ni at room temp. and atm. pressure yields VII. IV and VII were only slightly sol. in H2O (0.02 g./100 cc.) but were fairly sol. in dil. HCl; they are inferior in bacteriostatic activity to sulfathiazole.
- 22Bashford, V. G.; Wiggins, L. F. 82. Anhydrides of polyhydric alcohols. Part XIII. The amino-derivatives of 1 : 4–3 : 6-dianhydro-mannitol, -sorbitol, and -L-iditol, and their behaviour towards nitrous acid. J. Chem. Soc. 1950, 0, 371– 374, DOI: 10.1039/jr950000037122https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaG3cXisleiug%253D%253D&md5=3fd894628fd5232d88f72d0dbf4edb90Anhydrides of polyhydric alcohols. XIII. The amino derivatives of 1, 4:3, 6-dianhydromannitol, -sorbitol, and L-iditol and their behavior towards nitrous acidBashford, V. G.; Wiggins, L. F.Journal of the Chemical Society (1950), (), 371-4CODEN: JCSOA9; ISSN:0368-1769.cf. C.A. 43, 2939f; 44, 1O24d. 2, 5-Diamino-1, 4:3, 6-dianhydro-2, 5-didesoxysorbitol (I) (3.95 g.) in 25 cc. H2O, acidified with dil. HCl and treated with 4.5 g. NaNO2 in 25 cc. H2O, gives 1.27 g. 1,4:3,6-dianhydro-L-iditol (II), b0.05 115-30° (bath), [α]D 46.7° (Me2CO, c 1.9), characterized as the 2,5-bis(methylsulfonyl) deriv., m. 156.5° the small yield indicates that II was not the only product; the 2, 5-didesoxymannitol analog (III) of I also gives II; in the formation of II, deamination is accompanied by Walden inversion at C5 in the case of I and at both C2 and C5 in the case of III; 2,5-bis(p-tolylsulfonyl) deriv. (IV) of II, m. 90°, [α]D38.2°(CHCl3, c 2). IV(42g.) in 900 cc. MeOH, satd. at 0° with NH3, heated 30 hrs. at 160-70°, and the residue heated at 100° with 50 g. Ba(OH)2 in 400 cc. H2O, gives 2 g. 2,5-imino-1,4:3,6-dianhydro-2,5-didesoxy-D-mannitol (?) (V), m. 99-100°, [α]D 90.5° (CHCl3, c 2.65) [picrate, yellow, m. 219-20° (decompn.); HCl salt, m. 280-90° (decompn.); oxalate, m. 243°]; with NaNO2 in dil. AcOH, V forms an N-NO deriv., m. 121.5°, [α]D -323° (CHCl3, c 2.43). The conversion of I and III into II is paralleled by the epimerization of 2,4:3,5-dimethylene-D-manno- and -D-glucosaccharic acids (C.A. 38, 2631.8).
- 23Thiyagarajan, S.; Gootjes, L.; Vogelzang, W.; Wu, J.; van Haveren, J.; van Es, D. S. Chiral building blocks from biomass: 2,5-diamino-2,5-dideoxy-1,4-3,6-dianhydroiditol. Tetrahedron 2011, 67, 383– 389, DOI: 10.1016/j.tet.2010.11.03123https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXhsFCqs7rO&md5=dbb1b78c2a02975049665b0a4b5fca25Chiral building blocks from biomass: 2,5-diamino-2,5-dideoxy-1,4-3,6-dianhydroiditolThiyagarajan, Shanmugam; Gootjes, Linda; Vogelzang, Willem; Wu, Jing; van Haveren, Jacco; van Es, Daan S.Tetrahedron (2011), 67 (2), 383-389CODEN: TETRAB; ISSN:0040-4020. (Elsevier Ltd.)An efficient route towards the synthesis of 2,5-diamino-2,5-dideoxy-1,4-3,6-dianhydroiditol 4 has been developed resulting in significant improvements in both isolated yields and purity when compared to literature procedures. As a consequence, resin-grade 2,5-diamino-2,5-dideoxy-1,4-3,6-dianhydroiditol 4 has become available for lab. scale step-growth polymer synthesis. Addnl., an interesting renewable chiral 2-amino-2-deoxy-1,4-3,6-dianhydroiditol, has been isolated.
- 24Thiyagarajan, S.; Gootjes, L.; Vogelzang, W.; van Haveren, J.; Lutz, M.; van Es, D. S. Renewable rigid diamines: efficient, stereospecific synthesis of high purity isohexide diamines. ChemSusChem 2011, 4, 1823– 1829, DOI: 10.1002/cssc.20110039824https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXhsFeqt73O&md5=baeaf5ec2bbd64becd55551985f14c72Renewable Rigid Diamines: Efficient, Stereospecific Synthesis of High Purity Isohexide DiaminesThiyagarajan, Shanmugam; Gootjes, Linda; Vogelzang, Willem; van Haveren, Jacco; Lutz, Martin; van Es, Daan S.ChemSusChem (2011), 4 (12), 1823-1829CODEN: CHEMIZ; ISSN:1864-5631. (Wiley-VCH Verlag GmbH & Co. KGaA)We report an efficient three-step strategy for synthesizing rigid, chiral isohexide diamines derived from 1,4:3,6-dianhydrohexitols. These biobased chiral building blocks are presently the subject of several investigations (in our and several other groups) because of their application in high-performance biobased polymers, such as polyamides and polyurethanes. Among the three possible stereo-isomers, dideoxy-diamino isoidide and dideoxy-diamino isosorbide can be synthesized from isomannide and isosorbide resp. in high yield with abs. stereo control. Furthermore, by using this methodol. dideoxy-amino isomannide - a tricyclic adduct - was obtained starting from isoidide in high yield. Our improved synthetic route is a valuable advance towards meeting scale and purity demands for evaluating the properties of new biobased performance materials, which will benefit the development of these plastics.
- 25Kuszmann, J.; Medgyes, G. Synthesis and biological activity of 1,4:3,6-dianhydro-2,5-diazido-2,5-dideoxyhexitols. Carbohydr. Res. 1980, 85, 259– 269, DOI: 10.1016/s0008-6215(00)84675-325https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL3MXht1WrsL8%253D&md5=f0a8154a0a54251bcce73388a2dcc70cSynthesis and biological activity of 1,4:3,6-dianhydro-2,5-diazido-2,5-dideoxyhexitolsKuszmann, Janos; Medgyes, GaborCarbohydrate Research (1980), 85 (2), 259-69CODEN: CRBRAT; ISSN:0008-6215.Reaction of 1,4:3,6-dianhydro-2,5-di-O-mesyl- and -tosyl-D-mannitol with NaN3 afforded the 2,5-diazido-L-iditol deriv. The analogous D-glucitol isomer was obtained in a similar reaction starting from the corresponding D-glucitol derivs., and showed significant, hypnotic activity (no data). For establishing the structure-activity relationship, 1,4:3,6-dianhydro-2,5-diazido-2,5-dideoxy-L-mannitol (I), as well as its antipode (II), was synthesized, starting from D-mannitol. I was as effective as Doriden (3-ethyl-3-phenylglutarimide), a well known hypnotic drug. II and the bioisosteric 1(4),3(6)-dithio deriv. were, however, inactive.
- 26Bähn, S.; Imm, S.; Neubert, L.; Zhang, M.; Neumann, H.; Beller, M. The catalytic amination of alcohols. ChemCatChem 2011, 3, 1853– 1864, DOI: 10.1002/cctc.201100255There is no corresponding record for this reference.
- 27Pera-Titus, M.; Shi, F. Catalytic amination of biomass-based alcohols. ChemSusChem 2014, 7, 720– 722, DOI: 10.1002/cssc.20130109527https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXjtVGmsrY%253D&md5=bbb98a0b39689f5ecf960a64fa812d0bCatalytic Amination of Biomass-Based AlcoholsPera-Titus, Marc; Shi, FengChemSusChem (2014), 7 (3), 720-722CODEN: CHEMIZ; ISSN:1864-5631. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. Although alc. amination reactions were extensively studied since the early decades of the last century, the realization of sustainable development in the amine industry by implementing biomass-based alcs. as starting materials is still in its infancy. Catalytic systems based on Ru and Ir operating by means of the BH mechanism are currently available for performing such reactions. However, to reduce the prodn. costs, homogeneous systems based on cheaper metals operating by nucleophilic substitution, as well as supported metal nanoparticles (Ni, Co, Cu, Pd, Au) on low- alk. supports, are highly desired and new developments are expected to occur soon. With these catalysts in hand, new research areas for amination involving biomass-based alcs. with market opportunities are expected to be developed during this decade.
- 28Pingen, D.; Diebolt, O.; Vogt, D. Direct amination of bio-alcohols using ammonia. ChemCatChem 2013, 5, 2905– 2912, DOI: 10.1002/cctc.20130040728https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXhtVGktbfJ&md5=ad5a8a6a4ed29046a42e7eb4bfe6dff0Direct Amination of Bio-Alcohols Using AmmoniaPingen, Dennis; Diebolt, Olivier; Vogt, DieterChemCatChem (2013), 5 (10), 2905-2912CODEN: CHEMK3; ISSN:1867-3880. (Wiley-VCH Verlag GmbH & Co. KGaA)A slightly adapted catalyst system has been successfully applied in the direct amination of primary and secondary alcs. Moreover, the applicability to diols has been shown, giving high selectivity towards the primary diamines. It was found that the Ru/P ratio as well as the amt. of ammonia used are highly important in this system, esp. for higher substrate loadings. The catalyst was employed on a larger batch scale for the conversion of isomannide to the corresponding diamine. Addnl., it was shown that the catalyst is stable for at least six consecutive runs. No significant loss of activity and selectivity was obsd.
- 29Wright, L. W.; Brandner, J. D. Catalytic Isomerization of Polyhydric Alcohols.1 II. The Isomerization of Isosorbide to Isomannide and Isoidide. J. Org. Chem. 1964, 29, 2979– 2982, DOI: 10.1021/jo01033a04329https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaF2cXkvVCmtr0%253D&md5=1a81c31e7a0756aabc537182147720b1Catalytic isomerization of polyhydric alcohols. II. The isomerization of isosorbide to isomannide and isoidideWright, L. W.; Brandner, J. D.Journal of Organic Chemistry (1964), 29 (10), 2979-82CODEN: JOCEAH; ISSN:0022-3263.At 220-240° and 150 atm. H pressure the reversible interconversion of the 1,4:3,6-dianhydrohexitols of D-glucitol, D-mannitol, and L-iditol reaches a steady state after 2-6 h. in the presence of Ni-kieselguhr catalyst. At this time the approx. concns. are 57 % 1,4:3,6-dianhydro-L-iditol, 36% 1,4:3,6-dianhydro-D-glucitol; and 7% 1,4:3,6-dianhydro-D-mannitol. These figures are shown to be consistent with probability considerations, i.e., the relative amts. of the dianhydrohexitols are related to the probability of a given OH group being either exo or endo. Taking the steady-state mole fraction of 1,4:3,6-dianhydro-L-iditol as 0.57, it is calcd. that the probability of a OH being exo is 3 times the probability of its being endo. Calcn. of the mole fraction of the other 2 anhydrohexitols on the basis of these relative probabilities yields values in close agreement with those found exptl. The isomerization is strongly accelerated by increasing alky. of the catalyst-dianhydrohexitol slurry. Cf. CA 56, 11677c.
- 30Brandner, D.; Wright, L. W. Process for producing isoidide. U.S. Patent no. 3,023,223 A, 1962.There is no corresponding record for this reference.
- 31Schelwies, M.; Brinks, M.; Schaub, T.; Melder, J.-P.; Paciello, R.; Merger, M. Process for the homogeneously catalyzed amination of alcohols with ammonia in the presence of a complex catalyst which comprises nonanionic ligands. Patent no. WO 2014016241 A1 2014There is no corresponding record for this reference.
- 32Bahé, F.; Grand, L.; Cartier, E.; Jacolot, M.; Moebs-Sanchez, S.; Portinha, D.; Fleury, E.; Popowycz, F. Direct amination of isohexides via borrowing hydrogen methodology: regio- and stereoselective issues. Eur. J. Org. Chem. 2020, 2020, 599– 608, DOI: 10.1002/ejoc.20190166132https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXhs1ertr8%253D&md5=305cec7a4a4805b400ea26905f0549c3Direct Amination of Isohexides via Borrowing Hydrogen Methodology: Regio- and Stereoselective IssuesBahe, Florian; Grand, Lucie; Cartier, Elise; Jacolot, Maiwenn; Moebs-Sanchez, Sylvie; Portinha, Daniel; Fleury, Etienne; Popowycz, FlorenceEuropean Journal of Organic Chemistry (2020), 2020 (5), 599-608CODEN: EJOCFK; ISSN:1099-0690. (Wiley-VCH Verlag GmbH & Co. KGaA)The regio and diastereoselective direct mono or diamination of bio-based isohexides (isosorbide and isomannide) has been developed through borrowing hydrogen (BH) methodol. using a cooperative catalysis between an iridium complex and a Bronsted acid. The access to chiral amino-alc. (NH2-OH) and diamine (NH2-NH2), interesting optically pure bio-based monomers, was also proposed using BH strategy.
- 33Pfützenreuter, R.; Rose, M. Aqueous-phase amination of biogenic isohexides by using Ru/C as a solid catalyst. ChemCatChem 2016, 8, 251– 255, DOI: 10.1002/cctc.20150107733https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXhvFenurfE&md5=0207c7b9b607c92581601cfc22cfa8d3Aqueous-Phase Amination of Biogenic Isohexides by using Ru/C as a Solid CatalystPfuetzenreuter, Rebecca; Rose, MarcusChemCatChem (2016), 8 (1), 251-255CODEN: CHEMK3; ISSN:1867-3880. (Wiley-VCH Verlag GmbH & Co. KGaA)An aq. phase process for the amination of biogenic isohexides with ammonia using Ru/C as solid catalyst is reported. The products possess great potential for the prodn. of biogenic polyamides. From the diol substrates amino alcs. and diamines are derived as main products with significantly varying product selectivity depending on the substrates' stereochem. The catalytic reaction is performed at temps. of 140-180 °C. The Ru/C catalyst shows a high activity in the aq. phase, which is beneficial for the transformation of the polar biogenic substrates. Despite the basic conditions, the metal leaching is negligible and the catalyst can be recycled easily in batch operation and, thus, enables further process development due to the mild and scalable reaction conditions.
- 34Niemeier, J.; Engel, R. V.; Rose, M. Is water a suitable solvent for the catalytic amination of alcohols?. Green Chem. 2017, 19, 2839– 2845, DOI: 10.1039/c7gc00422b34https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXkvF2ku7k%253D&md5=ecdabfe39aabca9726addeebf734d262Is water a suitable solvent for the catalytic amination of alcohols?Niemeier, Johannes; Engel, Rebecca V.; Rose, MarcusGreen Chemistry (2017), 19 (12), 2839-2845CODEN: GRCHFJ; ISSN:1463-9262. (Royal Society of Chemistry)The catalytic conversion of biomass and biogenic platform chems. typically requires the use of solvents. Water is present already in the raw materials and in most cases a suitable solvent for the typically highly polar substrates. Hence, the development of novel catalytic routes for further processing would profit from the optimization of the reaction conditions in the aq. phase mainly for energetic reasons by avoiding the initial water sepn. Herein, we report the amination of biogenic alcs. in aq. solns. using solid Ru-based catalysts and ammonia as a reactant. The influence of different support materials and bimetallic catalysts is investigated for the amination of isomannide as a biogenic diol. Most importantly, the transferability of the reaction conditions to various other primary and secondary alcs. is successfully proved. Hence, water appears to be a suitable solvent for the sustainable prodn. of biogenic amines and offers great potential for further process development.
- 35Hu, H.; Ramzan, A.; Wischert, R.; Jerôme, F.; Michel, C.; de Olivera Vigier, K.; Pera-Titus, M. Pivotal role of H2 in the isomerisation of isosorbide over a Ru/C catalyst. Catal. Sci. Technol. 2021, 11, 7973– 7981, DOI: 10.1039/d1cy01709h35https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3MXit1KmtLjF&md5=2c5f64a0017eb471c2b29a740e4d49a7Pivotal role of H2 in the isomerization of isosorbide over a Ru/C catalystHu, H.; Ramzan, A.; Wischert, R.; Jerome, F.; Michel, C.; de Olivera Vigier, K.; Pera-Titus, M.Catalysis Science & Technology (2021), 11 (24), 7973-7981CODEN: CSTAGD; ISSN:2044-4753. (Royal Society of Chemistry)Isosorbide isomerisation is a known reaction that can proceed over Ru and Ni-based heterogeneous catalysts. As a rule, an exogenous H2 pressure (40-100 bar) is required, even though H2 does not participate stoichiometrically in the reaction. By marrying expts. with DFT computations, we ascribe the role of H2 in isosorbide isomerisation to a coverage effect on the catalyst surface. We demonstrate the possibility of conducting the reaction at a low H2 pressure either in the presence of an inert gas to increase H2 soly. in an underlying solvent or using 2-propanol as a hydrogen donor. This might benefit the economy and safety of a potential industrial process.
- 36Hausoul, P. J. C.; Negahdar, L.; Schute, K.; Palkovits, R. Unravelling the Ru-catalyzed hydrogenolysis of biomass-based polyols under neutral and acidic conditions. ChemSusChem 2015, 8, 3323– 3330, DOI: 10.1002/cssc.20150049336https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXhsVOms7fO&md5=f641f25935c7fa50f1829a2edf002e16Unravelling the Ru-Catalyzed Hydrogenolysis of Biomass-Based Polyols under Neutral and Acidic ConditionsHausoul, Peter J. C.; Negahdar, Leila; Schute, Kai; Palkovits, ReginaChemSusChem (2015), 8 (19), 3323-3330CODEN: CHEMIZ; ISSN:1864-5631. (Wiley-VCH Verlag GmbH & Co. KGaA)The aq. Ru/C-catalyzed hydrogenolysis of biomass-based polyols such as erythritol, xylitol, sorbitol, and cellobitol is studied under neutral and acidic conditions. For the first time, the complete product spectrum of C2-C6 polyols is identified and, based on a thorough anal. of the reaction mixts., a comprehensive reaction mechanism is proposed, which consists of (de)hydrogenation, epimerization, decarbonylation, and deoxygenation reactions. The data reveal that the Ru-catalyzed deoxygenation reaction is highly selective for the cleavage of terminal hydroxyl groups. Changing from neutral to acidic conditions suppresses decarbonylation, consequently increasing the selectivity towards deoxygenation.
- 37Wang, T.; Ibañez, J.; Wang, K.; Fang, L.; Sabbe, M.; Michel, C.; Paul, S.; Pera-Titus, M.; Sautet, P. Rational design of selective metal catalysts for alcohol amination with ammonia. Nat. Catal. 2019, 2, 773– 779, DOI: 10.1038/s41929-019-0327-237https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXhsFWlt7nF&md5=a01eb8c81f84ae1159a04fe6225f7d56Rational design of selective metal catalysts for alcohol amination with ammoniaWang, Tao; Ibanez, Javier; Wang, Kang; Fang, Lin; Sabbe, Maarten; Michel, Carine; Paul, Sebastien; Pera-Titus, Marc; Sautet, PhilippeNature Catalysis (2019), 2 (9), 773-779CODEN: NCAACP; ISSN:2520-1158. (Nature Research)The lack of selectivity for the direct amination of alcs. with ammonia (a modern and clean route for the synthesis of primary amines) is an unsolved problem. Here, we combine first-principles calcns., scaling relations, kinetic simulations and catalysis expts. to det. the key factors that govern the activity and selectivity of metal catalysts for this reaction. We show that the loss of selectivity towards primary amines is linked to a surface-mediated C-N bond coupling between two N-contg. intermediates: CH3NH and CH2NH. The barrier for this step is low enough to compete with the main surface hydrogenation reactions and it can be used as a descriptor for selectivity. The activity and selectivity maps (using the C and O adsorption energies as descriptors) were combined for the computational screening of 348 dil. bimetallic catalysts. Among the best theor. candidates, Co98.5Ag1.5 and Co98.5Ru1.5 (5 wt% Co) were identified exptl. to be the most promising catalysts.
- 38Honkala, K.; Hellman, A.; Remediakis, I. N.; Logadottir, A.; Carlsson, A.; Dahl, S.; Christensen, C. H.; Nørskov, J. K. Ammonia synthesis from first-principles calculations. Science 2005, 307, 555– 558, DOI: 10.1126/science.110643538https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2MXmslOjuw%253D%253D&md5=4c6066f84d200c9c587b6dd1dcd3b0bfAmmonia Synthesis from First-Principles CalculationsHonkala, K.; Hellman, A.; Remediakis, I. N.; Logadottir, A.; Carlsson, A.; Dahl, S.; Christensen, C. H.; Norskov, J. K.Science (Washington, DC, United States) (2005), 307 (5709), 555-558CODEN: SCIEAS; ISSN:0036-8075. (American Association for the Advancement of Science)The rate of ammonia synthesis over a nanoparticle ruthenium catalyst can be calcd. directly on the basis of a quantum chem. treatment of the problem using d. functional theory. We compared the results to measured rates over a ruthenium catalyst supported on magnesium aluminum spinel. When the size distribution of ruthenium particles measured by transmission electron microscopy was used as the link between the catalyst material and the theor. treatment, the calcd. rate was within a factor of 3 to 20 of the exptl. rate. This offers hope for computer-based methods in the search for catalysts.
- 39Lu, X.; Zhang, J.; Chen, W.-K.; Roldan, A. Kinetic and mechanistic analysis of NH3 decomposition on Ru(0001), Ru(111) and Ir(111) surfaces. Nanoscale Adv 2021, 3, 1624– 1632, DOI: 10.1039/d1na00015b39https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3MXjsVOgurg%253D&md5=e3fd50b5f5239cd7fe9e9fc3f182ed4bKinetic and mechanistic analysis of NH3 decomposition on Ru(0001), Ru(111) and Ir(111) surfacesLu, Xiuyuan; Zhang, Jing; Chen, Wen-Kai; Roldan, AlbertoNanoscale Advances (2021), 3 (6), 1624-1632CODEN: NAADAI; ISSN:2516-0230. (Royal Society of Chemistry)We investigated the catalytic NH3 decompn. on Ru and Ir metal surfaces using d. functional theory. The reaction mechanisms were unraveled on both metals, considering that, on the nano-scale, Ru particles may also present an fcc structure, hence, leading to three energy profiles. We implemented thermodn. and kinetic parameters obtained from DFT into microkinetic simulations. Batch reactor simulations suggest that hydrogen generation starts at 400 K, 425 K and 600 K on Ru(111), Ru(0001) and Ir(111) surfaces, resp., in excellent agreement with expts. During the reaction, the main surface species on Ru are NH, N and H, whereas on Ir(111), it is mainly NH. The rate-detg. step for all surfaces is the formation of mol. nitrogen. We also performed temp.-programmed reaction simulations and inspected the desorption spectra of N2 and H2 as a function of temp., which highlighted the importance of N coverage on the desorption rate.
- 40Benndorf, C.; Madey, T. E. Adsorption and orientation of NH3 on Ru(001). Surf. Sci. 1983, 135, 164– 183, DOI: 10.1016/0039-6028(83)90217-040https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL2cXnsVGisw%253D%253D&md5=f07ea305e45db23b65297650120d78d3Adsorption and orientation of ammonia on ruthenium(001)Benndorf, Carsten; Madey, Theodore E.Surface Science (1983), 135 (1-3), 164-83CODEN: SUSCAS; ISSN:0039-6028.The interaction of NH3 with clean Ru(001) surfaces was studied by using LEED ESDIAD (electron stimulated desorption ion angular distribution), TDS (thermal desorption spectroscopy), and work function changes (Δ.vphi.). Four different binding states (denoted as α1, α2, β, and γ) were detected with TDS. At low coverages, NH3 desorbs from the α1 state with a TDS peak max. at ∼310 K. The broadening of the TDS peaks and their shift to lower temp. with increasing NH3 coverage are related to repulsive lateral interactions between neighboring NH3 mols. At higher NH3 coverages (θNH3 ⪆ 0.15), a 2nd desorption peak (α2) develops at 180 K, accompanied by a (2 × 2) LEED structure. With further increase of NH3 exposure a sharp desorption peak (β state) is found at 140 K, an is interpreted as due to NH3 species desorbing from a 2nd adsorption layer. Finally a desorption peak due to multilayer adsorption (γ state) is found at 115 K. At low NH3 coverages (α1 state), a halo-like H+ ESDIAD pattern gives evidence of randomly oriented or freely rotating NH3 monomers, bounded via the N atoms to the surface with the H atoms pointing away from the surface. This orientation of NH3 is supported by work function measurements showing a linear decrease of Δ.vphi. in the α1 state. Structural information concerning the adsorption geometry of NH3 in the β state was obtained from LEED and ESDIAD. During the formation of the 2nd NH3 layer (β) a (2√3 × 2√3)R30° LEED pattern is obsd. and is accompanied by an ESDIAD pattern with a hexagonal outline. A structural model of the β-state bonding, in which 2nd layer NH3 mols. are bonded via 3-fold H bonds to the 1st layer NH3, is proposed.
- 41Carabineiro, S. A. C.; Matveev, A. V.; Gorodetskii, V. V.; Nieuwenhuys, B. E. Selective oxidation of ammonia over Ru(0001). Surf. Sci. 2004, 555, 83– 93, DOI: 10.1016/j.susc.2004.02.02241https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXislyltrs%253D&md5=e6e8139a87dd657fb1c6d2a6c54759d5Selective oxidation of ammonia over Ru(0 0 0 1)Carabineiro, S. A. C.; Matveev, A. V.; Gorodetskii, V. V.; Nieuwenhuys, B. E.Surface Science (2004), 555 (1-3), 83-93CODEN: SUSCAS; ISSN:0039-6028. (Elsevier Science B.V.)The decompn. and oxidn. of NH3 were studied on the Ru(0 0 0 1) surface in the temp. range from 150 up to 800 K. The results were compared to those found for Ir(1 1 0) and Ir(5 1 0). TDS results showed that most of the NH3 is dissociatively adsorbed between 150 and 300 K, with formation of H2 around 300 K and N2 between 600 and 800 K. N2 desorption shifts to lower temps. with increasing surface O coverage. The products of NH3 oxidn. obsd. were N2, H2O, and N2O. Formation of NO was not found. Inhibition of the reaction presumably by N species was obsd. until 450 and 670 K, depending on the NH3/O2 ratios. Above those temps. the reaction started as manifested by a decrease in the NH3 and O2 pressures and a simultaneous increase in the H2O, N2 and N2O pressures.
- 42Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A., Jr; Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, J. M.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, O. ̈.; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian 09; Gaussian, Inc.: Wallingford, CT, 2016.There is no corresponding record for this reference.
- 43Zhao, Y.; Truhlar, D. G. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: Two new functionals and systematic testing of four M06-class functionals and 12 other functionals. Theor. Chem. Acc. 2008, 120, 215– 241, DOI: 10.1007/s00214-007-0310-x43https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXltFyltbY%253D&md5=c31d6f319d7c7a45aa9b716220e4a422The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: two new functionals and systematic testing of four M06-class functionals and 12 other functionalsZhao, Yan; Truhlar, Donald G.Theoretical Chemistry Accounts (2008), 120 (1-3), 215-241CODEN: TCACFW; ISSN:1432-881X. (Springer GmbH)We present two new hybrid meta exchange-correlation functionals, called M06 and M06-2X. The M06 functional is parametrized including both transition metals and nonmetals, whereas the M06-2X functional is a high-nonlocality functional with double the amt. of nonlocal exchange (2X), and it is parametrized only for nonmetals. The functionals, along with the previously published M06-L local functional and the M06-HF full-Hartree-Fock functionals, constitute the M06 suite of complementary functionals. We assess these four functionals by comparing their performance to that of 12 other functionals and Hartree-Fock theory for 403 energetic data in 29 diverse databases, including ten databases for thermochem., four databases for kinetics, eight databases for noncovalent interactions, three databases for transition metal bonding, one database for metal atom excitation energies, and three databases for mol. excitation energies. We also illustrate the performance of these 17 methods for three databases contg. 40 bond lengths and for databases contg. 38 vibrational frequencies and 15 vibrational zero point energies. We recommend the M06-2X functional for applications involving main-group thermochem., kinetics, noncovalent interactions, and electronic excitation energies to valence and Rydberg states. We recommend the M06 functional for application in organometallic and inorganometallic chem. and for noncovalent interactions.
- 44Weigend, F.; Ahlrichs, R. Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys. Chem. Chem. Phys. 2005, 7, 3297– 3305, DOI: 10.1039/b508541a44https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2MXpsFWgu7o%253D&md5=a820fb6055c993b50c405ba0fc62b194Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracyWeigend, Florian; Ahlrichs, ReinhartPhysical Chemistry Chemical Physics (2005), 7 (18), 3297-3305CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)Gaussian basis sets of quadruple zeta valence quality for Rb-Rn are presented, as well as bases of split valence and triple zeta valence quality for H-Rn. The latter were obtained by (partly) modifying bases developed previously. A large set of more than 300 mols. representing (nearly) all elements-except lanthanides-in their common oxidn. states was used to assess the quality of the bases all across the periodic table. Quantities investigated were atomization energies, dipole moments and structure parameters for Hartree-Fock, d. functional theory and correlated methods, for which we had chosen Moller-Plesset perturbation theory as an example. Finally recommendations are given which type of basis set is used best for a certain level of theory and a desired quality of results.
- 45Weigend, F.; Ahlrichs, R. Accurate Coulomb-fitting basis sets for H to Rn. Phys. Chem. Chem. Phys. 2006, 8, 1057– 1065, DOI: 10.1039/b515623h45https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28Xhs12ntrc%253D&md5=314690393f1e21096541a317a80e563cAccurate Coulomb-fitting basis sets for H to RnWeigend, FlorianPhysical Chemistry Chemical Physics (2006), 8 (9), 1057-1065CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)A series of auxiliary basis sets to fit Coulomb potentials for the elements H to Rn (except lanthanides) is presented. For each element only one auxiliary basis set is needed to approx. Coulomb energies in conjunction with orbital basis sets of split valence, triple zeta valence and quadruple zeta valence quality with errors of typically below ca. 0.15 kJ mol-1 per atom; this was demonstrated in conjunction with the recently developed orbital basis sets of types def2-SV(P), def2-TZVP and def2-QZVPP for a large set of small mols. representing (nearly) each element in all of its common oxidn. states. These auxiliary bases are slightly more than three times larger than orbital bases of split valence quality. Compared to non-approximated treatments, computation times for the Coulomb part are reduced by a factor of ca. 8 for def2-SV(P) orbital bases, ca. 25 for def2-TZVP and ca. 100 for def2-QZVPP orbital bases.
- 46Luchini, G.; Alegre-Requena, J. V.; Funes-Ardoiz, I.; Paton, R. S. GoodVibes: automated thermochemistry for heterogeneous computational chemistry data; F1000 Research Ltd, 2020, Vol 9.There is no corresponding record for this reference.
- 47Alecu, I. M.; Zheng, J.; Zhao, Y.; Truhlar, D. G. Computational thermochemistry: scale factor databases and scale factors for vibrational frequencies obtained from electronic model chemistries. J. Chem. Theory Comput. 2010, 6, 2872– 2887, DOI: 10.1021/ct100326h47https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXhtVGrsb7E&md5=f0104fdf97972d89e7cb60e634b19c4bComputational Thermochemistry: Scale Factor Databases and Scale Factors for Vibrational Frequencies Obtained from Electronic Model ChemistriesAlecu, I. M.; Zheng, Jingjing; Zhao, Yan; Truhlar, Donald G.Journal of Chemical Theory and Computation (2010), 6 (9), 2872-2887CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Optimized scale factors for calcg. vibrational harmonic and fundamental frequencies and zero-point energies have been detd. for 145 electronic model chemistries, including 119 based on approx. functionals depending on occupied orbitals, 19 based on single-level wave function theory, three based on the neglect-of-diat.-differential-overlap, two based on doubly hybrid d. functional theory, and two based on multicoefficient correlation methods. Forty of the scale factors are obtained from large databases, which are also used to derive two universal scale factor ratios that can be used to interconvert between scale factors optimized for various properties, enabling the derivation of three key scale factors at the effort of optimizing only one of them. A reduced scale factor optimization model is formulated in order to further reduce the cost of optimizing scale factors, and the reduced model is illustrated by using it to obtain 105 addnl. scale factors. Using root-mean-square errors from the values in the large databases, we find that scaling reduces errors in zero-point energies by a factor of 2.3 and errors in fundamental vibrational frequencies by a factor of 3.0, but it reduces errors in harmonic vibrational frequencies by only a factor of 1.3. It is shown that, upon scaling, the balanced multicoefficient correlation method based on coupled cluster theory with single and double excitations (BMC-CCSD) can lead to very accurate predictions of vibrational frequencies. With a polarized, minimally augmented basis set, the d. functionals with zero-point energy scale factors closest to unity are MPWLYP1M (1.009), τHCTHhyb (0.989), BB95 (1.012), BLYP (1.013), BP86 (1.014), B3LYP (0.986), MPW3LYP (0.986), and VSXC (0.986).
- 48Ribeiro, R. F.; Marenich, A. V.; Cramer, C. J.; Truhlar, D. G. Use of solution-phase vibrational frequencies in continuum models for the free energy of solvation. J. Phys. Chem. B 2011, 115, 14556– 14562, DOI: 10.1021/jp205508z48https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXhsFSjtr3O&md5=3a164fbab7255d92e1099064e7f72261Use of Solution-Phase Vibrational Frequencies in Continuum Models for the Free Energy of SolvationRibeiro, Raphael F.; Marenich, Aleksandr V.; Cramer, Christopher J.; Truhlar, Donald G.Journal of Physical Chemistry B (2011), 115 (49), 14556-14562CODEN: JPCBFK; ISSN:1520-5207. (American Chemical Society)We find that vibrational contributions to a solute's free energy are in general insensitive to whether the solute vibrational frequencies are computed in the gas phase or in soln. In most cases, the difference is smaller than the intrinsic error in solvation free energies assocd. with the continuum approxn. to solvation modeling, although care must be taken to avoid spurious results assocd. with limitations in the quantum-mech. harmonic-oscillator approxn. for very low-frequency mol. vibrations. We compute solute vibrational partition functions in aq. and carbon tetrachloride soln. and compare them to gas-phase mol. partition functions computed with the same level of theory and the same quasiharmonic approxn. for the diverse and extensive set of mols. and ions included in the training set of the SMD continuum solvation model, and we find mean unsigned differences in vibrational contributions to the solute free energy of only about 0.2 kcal/mol. On the basis of these results and a review of the theory, we conclude, in contrast to previous work, that using partition functions computed for mols. optimized in soln. is a correct and useful approach for averaging over solute degrees of freedom when computing free energies of solutes in soln., and it is moreover recommended for cases where liq. and gas-phase solute structures differ appreciably or when stationary points present in liq. soln. do not exist in the gas phase, for which we provide some examples. When gas-phase and soln.-phase geometries and frequencies are similar, the use of gas-phase geometries and frequencies is a useful approxn.
- 49Li, Y.-P.; Gomes, J.; Mallikarjun Sharada, S.; Bell, A. T.; Head-Gordon, M. Improved force-field parameters for QM/MM simulations of the energies of adsorption for molecules in zeolites and a free rotor correction to the rigid rotor harmonic oscillator model for adsorption enthalpies. J. Phys. Chem. C 2015, 119, 1840– 1850, DOI: 10.1021/jp509921r49https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXitFehsbbJ&md5=d5fd7f41d9a44592826c59eea27f73c1Improved Force-Field Parameters for QM/MM Simulations of the Energies of Adsorption for Molecules in Zeolites and a Free Rotor Correction to the Rigid Rotor Harmonic Oscillator Model for Adsorption EnthalpiesLi, Yi-Pei; Gomes, Joseph; Mallikarjun Sharada, Shaama; Bell, Alexis T.; Head-Gordon, MartinJournal of Physical Chemistry C (2015), 119 (4), 1840-1850CODEN: JPCCCK; ISSN:1932-7447. (American Chemical Society)Quantum mechanics/mol. mechanics (QM/MM) simulations provide an efficient avenue for studying reactions catalyzed in zeolite systems; however, the accuracy of such calcns. is highly dependent on the zeolite MM parameters used. Previously reported parameters (P1), which were chosen to minimize the root mean square (RMS) deviations of adsorption energies compared with full QM ωB97X-D/6-31+G** adsorption energies, are shown to overestimate binding energies compared with exptl. values, particularly for larger substrates. To address this issue, a new parameter set (P2) is derived by rescaling the previously reported characteristic energies of the Lennard-Jones potential in P1. The accuracy of the thermal correction for adsorption enthalpies detd. by the rigid rotor-harmonic oscillator approxn. (RRHO) is examd. and shown to be improved by treating low-lying vibrational modes as free translational and rotational modes via a quasi-RRHO model. With P2 and quasi-RRHO, adsorption energies calcd. with QM/MM agree with exptl. values with an RMS error of 1.8 kcal/mol for both nonpolar and polar mols. adsorbed in MFI, H-MFI, and H-BEA. By contrast, the RMS error for the same test sets obtained using parameter set P1 is 8.3 kcal/mol. Glucose-fructose isomerization catalyzed by Sn-BEA is taken as an example to demonstrate that improved values for apparent activation energies can be obtained using the methodol. reported here. With parameter set P2, the apparent activation energy calcd. with QM/MM reproduces the exptl. value to within 1 kcal/mol. By contrast, using parameter set P1, the error is -12.9 kcal/mol.
- 50Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. J. Phys. Rev. B 1996, 54, 11169– 11186, DOI: 10.1103/physrevb.54.1116950https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK28Xms1Whu7Y%253D&md5=9c8f6f298fe5ffe37c2589d3f970a697Efficient 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.
- 51Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865– 3868, DOI: 10.1103/physrevlett.77.386551https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK28XmsVCgsbs%253D&md5=55943538406ee74f93aabdf882cd4630Generalized 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.
- 52Steinmann, S.; Corminboeuf, C. A generalized-gradient approximation exchange hole model for dispersion coefficients. J. Chem. Phys. 2011, 134, 044117, DOI: 10.1063/1.354598552https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXht1Giur4%253D&md5=c28723818938f9c1326bcb551835d562A generalized-gradient approximation exchange hole model for dispersion coefficientsSteinmann, Stephan N.; Corminboeuf, ClemenceJournal of Chemical Physics (2011), 134 (4), 044117/1-044117/5CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)A simple method for computing accurate d.-dependent dispersion coeffs. is presented. The dispersion coeffs. are modeled by a generalized gradient-type approxn. to Becke and Johnson's exchange hole dipole moment formalism. Our most cost-effective variant, based on a disjoint description of atoms in a mol., gives mean abs. errors in the C6 coeffs. for 90 complexes below 10%. The inclusion of the missing long-range van der Waals interactions in d. functionals using the derived coeffs. in a pair wise correction leads to highly accurate typical noncovalent interaction energies. (c) 2011 American Institute of Physics.
- 53Blöchl, P. E. Projector augmented-wave method. Phys. Rev. B 1994, 50, 17953– 17979, DOI: 10.1103/physrevb.50.1795353https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC2sfjslSntA%253D%253D&md5=1853d67af808af2edab58beaab5d3051Projector augmented-wave methodBlochlPhysical review. B, Condensed matter (1994), 50 (24), 17953-17979 ISSN:0163-1829.There is no expanded citation for this reference.
- 54Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758– 1775, DOI: 10.1103/physrevb.59.175854https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXkt12nug%253D%253D&md5=78a73e92a93f995982fc481715729b14From 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.
- 55Hermes, E. D.; Janes, A. N.; Schmidt, J. R. M. Micki: A python-based object-oriented microkinetic modeling code. J. Chem. Phys. 2019, 151, 014112, DOI: 10.1063/1.510911655https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXhtlahs7nK&md5=97a4449d51c0d1a40a261e77ea73f76aMicki: A python-based object-oriented microkinetic modeling codeHermes, Eric D.; Janes, Aurora N.; Schmidt, J. R.Journal of Chemical Physics (2019), 151 (1), 014112/1-014112/13CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)We have developed a flexible, general-purpose microkinetic modeling code, Micki, to analyze complex, heterogeneously catalyzed chem. reactions based upon first-principles calcns. This Python-based code is modular and object oriented, framing the development of microkinetic models in familiar chem. terms. We also present novel approaches, incorporated into Micki, to describe diffusion limited reactions, multidentate bindings, thermodynamically consistent lateral interactions, and Bronsted-Evans-Polanyi ests. of changes in barrier heights. Micki has built-in modules for subsequent anal. of microkinetic models, including degree of rate control and rate order. As a demonstration of the power and flexibility of the code, we build a microkinetic model for the water-gas shift reaction and compare to previously published exptl. results and microkinetic models, showing that Micki can quant. reproduce exptl. turnover frequencies with minimal empirical optimization. (c) 2019 American Institute of Physics.
Supporting Information
Supporting Information
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acssuschemeng.2c07501.
Coordinates for DFT calculations (ZIP)
Blank mass spectra of the MS system; mass spectra of AS2 and DAS; mass spectra of the reaction mixture after amination of isomannide IM under 30 bar H2; possible byproducts from IM amination; reductive amination of endo-OH ketone mixtures; optimized structures for different adsorption modes of aminoalcohols through either −OH or −NH2 functional groups; reaction profile corresponding to the dehydrogenation of NH3 to form NHx species on Ru(0001); GC methods A and B used for the analysis of product mixtures in the amination of IM; representative GC chromatograms obtained for reactant and product analysis using GC methods A and B; GC method C used for the analysis of product mixtures in the oxidation of IM and IS; representative GC chromatograms obtained for reactant and product analysis using GC methods A and C; steady-state coverage of surface species and empty catalytic sites as a function of temperature at a H2/NH3 ratio of 1.0; evolution of N2 gas as a function of temperature for H2/NH3 molar ratio of 1.0; and a table listing substrate conversion and product yield for figures in the main text (PDF)
Terms & Conditions
Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.