Simple, Chemoselective Hydrogenation with Thermodynamic Stereocontrol (2024)

  • Journal List
  • ACS AuthorChoice
  • PMC3951266

As a library, NLM provides access to scientific literature. Inclusion in an NLM database does not imply endorsem*nt of, or agreement with, the contents by NLM or the National Institutes of Health.
Learn more: PMC Disclaimer | PMC Copyright Notice

Simple,Chemoselective Hydrogenation with ThermodynamicStereocontrol (1)

J Am Chem Soc. 2014 Jan 29; 136(4): 1300–1303.

Published online 2014 Jan 15. doi:10.1021/ja412342g

PMCID: PMC3951266

NIHMSID: NIHMS557222

PMID: 24428640

Kotaro Iwasaki, Kanny K. Wan, Alberto Oppedisano, Steven W. M. Crossley, and Ryan A. Shenvi*

Author information Article notes Copyright and License information PMC Disclaimer

Associated Data

Supplementary Materials

Abstract

Simple,Chemoselective Hydrogenation with ThermodynamicStereocontrol (2)

Few methods permit the hydrogenationof alkenes to a thermodynamicallyfavored configuration when steric effects dictate the alternativetrajectory of hydrogen delivery. Dissolving metal reduction achievesthis control, but with extremely low functional group tolerance. Herewe demonstrate a catalytic hydrogenation of alkenes that affords thethermodynamic alkane products with remarkably broad functional groupcompatibility and rapid reaction rates at standard temperature andpressure.

Hydrogenation of alkenes isamong the cardinal reactions available to synthetic chemists. Appliedretrosynthetically, this transform adds points of unsaturation tocarbon skeletons, which can be used to dissect complex bond networksinto simple fragments.1 However, a long-standingchallenge in complex molecule synthesis is hydrogenation of alkenesto a thermodynamically favored configuration2 when steric constraints of the substrate favor hydrogenation toa nonthermodynamic (kinetic) alkane product (see 1cis-2, Figure ​Figure1a).1a). Dissolvingmetal reduction provides a means to achieve thermodynamic controland can reduce conjugated3 or electron-pooralkenes4 at low temperature (see 1trans-2), but requiresambient or elevated temperature to reduce electron-neutral alkenes.5 These latter substrates are therefore seldomemployed because chemoselectivity is unsatisfactory: most other functionalgroups are reduced preferentially to electron-neutral alkenes. Thisproblem is especially evident in chemical syntheses of terpenoid secondarymetabolites, which include the FDA approved steroid, taxoid, artemisinin,and ingenoid classes. Thus, many terpenes that contain equatorialmethyl groups, for instance 35,have proven difficult to access, since hydrogenation of the correspondingexomethylene using standard methods yields either the wrong epimeror an equimolar mixture of two epimers (Figure ​(Figure11b).68 The origin of the poor chemoselectivity associated with dissolvingmetal reduction is the low reduction potential of electron neutralalkenes (Ered < –3 V,Pb cathode),9 which form high-energy radicalanions (67, Figure ​Figure1c)1c) prior to protonation to lower energy tertiary radicals(8) and further reduction, protonation to alkanes (9). We thought that circumvention of radical anion 7 via direct hydrogen atom transfer (HAT)10 might increase chemoselectivity but lead to the same stereochemicalpreferences as dissolving metal reduction. Here we show that manganeseand cobalt catalysts can effect this stepwise radical hydrogenationof electron-neutral alkenes and exhibit the same stereochemical preferencesas dissolving metals but spare a variety of reactive functional groupsthat are normally reduced.

Open in a separate window

Figure 1

Kinetic versus thermodynamic hydrogenation.(a) Example of stereodivergenthydrogenation; (b) examples where kinetic hydrogenation yields theincorrect stereoisomer; (c) poor chemoselectivity of dissolving metalreduction might be circumvented by HAT hydrogenation.

As a representative model system for the terpenesshown in Figure ​Figure1b1b and as a starting pointto probe the concept andconsequences of HAT hydrogenation, we chose 4-tert-butylmethylenecyclohexane 10, since extensive dataon its hydrogenation are available (Table 1).11,12 Whereas hydrogenation of 10 with Wilkinson’s12 or Adam’scatalyst13 delivers primarily 12 (cis), bearing an axial methyl, and diimide reductiongives a 1:1 mixture,14 dissolving metalreduction instead provides high selectivity for 11 (95:5).15 It seemed reasonable that hydrogenation viaa tertiary carbon radical might lead to high selectivity for an equatorialmethyl group, since methyl-substituted cyclohexyl radicals are knownto trap tributyltin deuteride with high axial selectivity for C–Dbond formation, owing to the nonplanar ground state16 and transition state pyramidalization of tertiary radicals.17 We considered the redox hydration conditionsof Mukaiyama18 as a good starting pointfor an iterative HAT hydrogenation of alkenes since tertiary radicalsare readily generated from electron-neutral alkenes, and the cobaltor manganese catalysts19 are inexpensiveand air-stable. However, the oxidizing conditions required for thistransformation appeared to be an ostensible barrier to implementationof a reductive reaction.20 Even so, Carreirademonstrated that tert-butyl hydroperoxide can serveas a replacement activator and/or reoxidant in the context of cobalt-catalyzedhydroazidation19 and hydrocyanation ofalkenes.21 Similarly, we found that ifTBHP is added in stoichiometric quantities to a mixture of alkene 10, phenylsilane, and a manganese catalyst in the absenceof heteroatom radical traps, the reaction rapidly produces 12 (trans) with high selectivity, albeit in poor yield.Higher conversion to hydrogenated products is highly ligand and solventdependent; the dipivaloyl-methane (dpm) ligand and isopropanol solventwere found to be optimal.22 Most appealingly,the experimental procedure is simple, conversion occurs usually within1 h, and no hydrogen atmosphere is required. Nonanhydrous conditionsare tolerated, however excess water does inhibit conversion (10 equiv= 5 M water in i-PrOH, 33% conversion at 1 h; see Supporting Information (SI) for an optimizationtable and corresponding observations). Co(dpm)2 inducesequally high levels of stereoselectivity and also allows nonpolarsolvents and hydrophobic substrates to be used (vide infra), however Mn(dpm)3-catalyzed reactions are generallyhigher yielding.

Table 1

Comparison to Existing Methods

Simple,Chemoselective Hydrogenation with ThermodynamicStereocontrol (4)

Open in a separate window

conditionsyield11:12
hom*ogeneous:5mol%RhCl(PPh3)3,1atmH2,PhH,18°C100%32:68
heterogeneous: 9mol% PtO2, 2 atm H2, AcOH100%21:79
pericyclic:N2H4, O2, EtOH, 55°CND49:51
dissolving metal: Li0, EDA, 35°CND95:5
radical: 10mol% Mn(dpm)3, 1.0 equiv PhSiH3, 1.5 equiv TBHP, i-PrOH (0.5 M), 22 °C, 1h86%84:16
or: 10mol% Co(dpm)2, 1.0 equiv PhSiH3, 1.5 equiv TBHP, i-PrOH (0.5 M), 22 °C, 1h69%86:14

Open in a separate window

As illustrated by Table 2a, trans-selective hydrogenation of cyclohexenes is difficultto achieveby standard methods. For instance 13a and b afford poor stereoselectivity using Pd/C catalysis, and only marginallybetter ratios using Crabtree’s catalyst, due to the poor directivityafforded by sterically encumbered substituents.23 However, using HAT hydrogenation, where directivity isirrelevant, trans-substitution with high stereoselectivityis favored in both cases.

As a result, diversely substitutedcyclohexenes now can be hydrogenatedto the thermodynamically preferred trans-isomers(Table 2b) in the absence of any directinggroups.23b Remarkably, chloro- and bromoalkenes,which generally do not tolerate dissolving metal conditions and arechallenging substrates even with standard hydrogenation methods, areefficiently reduced to equatorial halo-cyclohexanes 15 and 16. Electron-releasing siloxy (enolsilane) andacetamide (enamide) groups are also efficiently reduced to the trans-isomers 17 and 18, illustratingthe electronic flexibility of the method.

Table 2

InitialSurvey of Method’sUtility

Simple,Chemoselective Hydrogenation with ThermodynamicStereocontrol (5)

Open in a separate window

Simple,Chemoselective Hydrogenation with ThermodynamicStereocontrol (6)

Open in a separate window

As suggested by substrates 1518, the functional group compatibility of ourmethod is significantlybetter than dissolving metal reduction and a fuller scope is illustratedby Table 3. For instance, not only alcohols(19) but also their corresponding alkyliodides (20) are viable substrates for this transformation. Selectivityagainst the reduction of aromatic rings is excellent, and thereforecarboxybenzyl (Cbz) groups (21) are spared from reduction,as are phenyl ethers (22), aryl chlorides and fluorides(23), aryl bromides (24), and aryl iodides(25), although some deiodination was observed by GCMS(∼5%). Trifluoromethyl ethers (26), electron-richarenes (27), phenyl thioethers (28), trifluoromethylarenes(29), and heterocycles like imidazole 30 are all well tolerated, although some rate decrease is observedwith this last entry, likely due to equilibrium coordination/deactivationof the catalyst.

Table 3

Functional Group Tolerance

Simple,Chemoselective Hydrogenation with ThermodynamicStereocontrol (7)

Open in a separate window

Simple,Chemoselective Hydrogenation with ThermodynamicStereocontrol (8)

Open in a separate window

Weinreb amides, which suffer N–O bond cleavageunder dissolvingmetal conditions,24 do not react competitivelywith the reduction of trisubstituted alkenes (32, Table 4). Thioesters undergo facile reduction to aldehydesusing tributylstannyl radical reduction (Bu3SnH, AIBN)or palladium catalysis (Pd/C, Et3SiH),25 but using our method, saturated thioester 33 is produced and no aldehyde or alcohol is observed. When the alkeneis allylic to a potentially labile C-heteroatom bond, simple to complexheterocycles (3436) are toleratedand no scission of the allylic bond is observed, which might argueagainst the intermediacy of a carbon–metal bond.18,19 No Minisci addition products were observed using these substrates.Interestingly, β-ionone can be selectively reduced at the α,β-positions(37),20b in contrast to theselectivity observed using Mukaiyama’s hydration.20 Unsaturated thiols, aldehydes, and allylic alcohols(3840) are also chemoselectivelyreduced; the thiol is oxidized in situ to the disulfidewhich can be cleaved on workup.

Table 4

Diverse UnsaturatedSubstrates areReduced

Simple,Chemoselective Hydrogenation with ThermodynamicStereocontrol (9)

Open in a separate window

aUsing Co(dpm)2 conditions,

bUsing Mn(dpm)3 conditions.

Polycyclic systems can alsobe predictably hydrogenated to thethermodynamically stable diastereomer in preference to the normallyobserved kinetic stereochemistry (Scheme 1).For instance, Δ9,10-octalin (41) ispreferentially hydrogenated with iridium catalysis to cis-decalin (42a) via syn-hydrogenation of the alkene.26 However, using HAT hydrogenation, a formal anti-additionof hydrogen is observed to produce the lower-energy trans-decalin (42b) with nearly the same selectivity as dissolvingmetal conditions.5 Similarly, 1,2-dimethylcyclohexene (43) is hydrogenated to the cis-stereoisomer 44a using iridium catalysis,27 whereas our method produces trans-1,2-dimethylcyclohexane (44b), albeit with lower selectivity.Nevertheless, this stereochemical dichotomy between existing hydrogenationmethods and the title reaction can be most vividly illustrated usingproblems already encountered in terpene syntheses. For example, enroute to the putative structure of a sesquiterpene isolated from Cistus creticus, Katerinopoulos could not directly accessa targeted trans-decalone framework since hydrogenationof their intermediate ketone 45 using heterogeneous catalysisproduced cis-decalone 46a with highselectivity.28 In contrast, HAT hydrogenationcan directly access this thermodynamically preferred but kineticallydisfavored configuration (46b). Additionally, the terpene-derivedpetroleum biomarker drimane (48b) could not be directlyaccessed via hydrogenation of drimene 47, since use ofAdam’s catalyst delivers the kinetically favored axial methylsubstituent in 48a, and a four-step work-around was devisedinstead.29 In contrast, our method directlyyields drimane (48b), which bears the thermodynamicallyfavored equatorial methyl.

Open in a separate window

Scheme 1

Examples of Divergent Stereocontrol

41 contained 16%of the 1,9-isomer.

43 contained 26% of the 1,6-isomer.

An unusual aspect of this hydrogenation is its general disregardfor alkene substitution or electronic patterning (Scheme 2a). Unlike dissolving metal reduction, which exhibitsprofound rate acceleration when the alkene is conjugated to an electronacceptor,5 various substitution patternsare readily reduced by our HAT method with little influence by directlyattached functionality (see SI for competitionexperiments). These minor effects of electron-modulating groups mayindicate a direct hydrogen atom transfer to generate a carbon-centeredradical,30b rather than the intermediacyof a carbon–metal bond en route to carbon–metal bondhom*olysis, as often proposed.18,19,22 Based on competition experiments, some trends are observed: increasedsubstitution decreases the rate of consumption; electron-withdrawinggroups have a minor accelerating effect on reaction rate; and electron-donatinggroups are weakly deactivating. Consequently, simple reductive cyclizationsof polyenes are possible (Scheme 2b).30 For instance, diene 49 is readilycyclized to cyclopentane 50, which bears vicinal all-carbonquaternary centers, highlighting the ability of carbon-centered radicalsto overcome severe steric clash via early transition states.31 This preliminary example paves the way for ageneral approach to reductive mono- and polycyclizations of unconjugatedpolyenes and a “traceless” approach to fully saturatedmolecules like terpenes.32

Open in a separate window

Scheme 2

ObservedReactivity Trends and Cyclizations

The demonstrated utility of this new method for the thermodynamichydrogenation of alkenes illustrates its potential to solve currentand future problems in complex molecule synthesis, especially thereduction of halogenated alkenes (see Table 2b).33 Given the frequency with which hydrogenationsare employed and their breadth of applications in chemistry, a newtool is advantageous. We expect that the experimental ease, broadscope, and orthogonal stereoselectivity of HAT hydrogenation willlead to its extensive use.

Acknowledgments

Financial supportfor this work was provided by the NIH (GM104180),the NSF GRFP (K.K.W.; DGE-1346837), and the Italian Ministry for Educationand Research (fellowship to A.O.). We thank Professors Dale L. Boger,Phil S. Baran, and Donna G. Blackmond for helpful conversations. Weare grateful to the Scripps Research Institute, Eli Lilly, BoehringerIngelheim, Amgen, and the Baxter Foundation for additional financialsupport.

Funding Statement

National Institutes of Health, United States

Supporting Information Available

Experimental procedures andspectroscopic data. This material is available free of charge viathe Internet at http://pubs.acs.org.

Author Present Address

Department of Chemistry, Tohoku University, 6-3 Aoba-ku, Sendai,980-8578, Japan.

Author Contributions

Theseauthors contributed equally.

Notes

Theauthors declare nocompeting financial interest.

Supplementary Material

ja412342g_si_001.pdf(12M, pdf)

References

  • Forexample:Corey E. J.; Cheng X.-M.. The Logic of Chemical Synthesis; Wiley: New York, 1995; pp 47–57. [Google Scholar]
  • Barton D. H. R.; Robinson C. H.J. Chem. Soc.1954, 3045. [Google Scholar]
  • Johnson W. S.; Bannister B.; Bloom B. M.; Kemp A. D.; Pappo R.; Rogier E. R.; Szmuszkovicz J.J. Am. Chem. Soc.1953, 75, 2275. [Google Scholar]
  • Stork G.; Darling S. D.J. Am. Chem. Soc.1960, 82, 1512. [Google Scholar]
  • Whitesides G. M.; Ehmann W. J.J. Org. Chem.1970, 35, 3565. [Google Scholar]
  • Takahashi S.; Kusumi T.; Kakisawa H.Chem. Lett.1979, 515. [Google Scholar]
  • Alexander R.; Kagi R.; Noble R.J. Chem. Soc., Chem.Comm.1983, 226. [Google Scholar]
  • Justicia J.; Rosales A.; Buñuel E.; Oller-López J. L.; Valdivia M.; Haïdour A.; Oltra J. E.; Barrero A. F.; Cárdenas D. J.; Cuerva J. M.Chem.—Eur. J.2004, 10, 1778. [PubMed] [Google Scholar]
  • Budnikova Y. G.; Krasnov S. A. In Developments inElectrochemistry;Chun J. H., Ed.; InTech: Rijeka, 2012; Chapter 5, pp 101–124. [Google Scholar]
  • Seminal study of ahydrogenation where HAT is implicated:Sweany R. L.; Halpern J.J. Am. Chem. Soc.1977, 99, 8335. [Google Scholar]
  • Imaizumi S.; Murayama H.; Ishiyama J.; Senda Y.Bull. Chem. Soc. Jpn.1985, 58, 1071. [Google Scholar]
  • Mitchell T. R. B.J. Chem. Soc.B1970, 823. [Google Scholar]
  • Siegel S.; Dmuchovsky B.J. Am. Chem. Soc.1962, 84, 3132. [Google Scholar]
  • vanTamelen E. E.; Timmons R. J.J. Am. Chem. Soc.1962, 84, 1067. [Google Scholar]
  • Ficini J.; Francillette J.; Touzin A. M.J. Chem. Research (S)1979, 150. [Google Scholar]
  • Lloyd R. V.; Williams R. V.J. Phys. Chem.1985, 89, 5379. [Google Scholar]
  • Damm W.; Giese B.; Hartung J.; Hasskerl T.; Houk K. N.; Hüter O.; Zipse H.J. Am. Chem. Soc.1992, 114, 4067. [Google Scholar]
  • Review and referencesto early work:Mukaiyama T.; Yamada T.Bull. Chem. Soc. Jpn.1995, 68, 17. [Google Scholar]
  • Waser J.; Gaspar B.; Nambu H.; Carreira E. M.J. Am. Chem. Soc.2006, 128, 11693. [PubMed] [Google Scholar]
  • a. Magnus P.; Payne A. H.; Waring M. J.; Scott D. A.; Lynch V.Tetrahedron Lett.2000, 41, 9725. [Google Scholar]
    b. Magnus P.; Waring M. J.; Scott D. A.Tetrahedron Lett.2000, 41, 9731.InRef (20b), the catalystturns overby protonation of a manganese enolate. [Google Scholar]
  • Carreira’s pioneering work in thisarea:
    a. Gaspar B.; Waser J.; Carreira E. M.Org. Syn.2010, 87, 88. [Google Scholar]
    b. Gaspar B.; Carreira E. M.J. Am. Chem. Soc.2009, 131, 13214. [PubMed] [Google Scholar]
    c. Gaspar B.; Carreira E. M.Angew. Chem., Int. Ed.2008, 47, 5758. [PubMed] [Google Scholar]
    d. Gaspar B.; Waser J.; Carreira E. M.Synthesis2007, 3839. [Google Scholar]
    e. Gaspar B.; Carreira E. M.Angew. Chem.,Int. Ed.2007, 46, 4519. [PubMed] [Google Scholar]
    f. Waser J.; González-Gómez J. C.; Nambu H.; Huber P.; Carreira E. M.Org. Lett.2005, 7, 4249. [PubMed] [Google Scholar]
    g. Waser J.; Nambu H.; Carreira E. M.J. Am. Chem. Soc.2005, 127, 8294. [PubMed] [Google Scholar]
    h. Waser J.; Carreira E. M.Angew. Chem., Int. Ed.2004, 43, 4099. [PubMed] [Google Scholar]
    i. Waser J.; Carreira E. M.J. Am. Chem. Soc.2004, 126, 5676. [PubMed] [Google Scholar]
  • Inoki S.; Kato K.; Isayama S.; Mukaiyama T.Chem. Lett.1990, 1869. [Google Scholar]
  • a. Crabtree R. H.; Davis M. W.J. Org. Chem.1986, 51, 2655. [Google Scholar]
    b. Excellent reviewon directed hom*ogenous hydrogenation:Brown J. M.Angew. Chem., Int. Ed.1987, 26, 190. [Google Scholar]
  • Taillier C.; Bellosta V.; Meyer C.; Cossy J.Org.Lett.2004, 6, 2145. [PubMed] [Google Scholar]
  • f*ckuyma T.; Tokuyama H.Ald. Acta2004, 37, 87. [Google Scholar]
  • Weitkamp A. W.J. Catal.1966, 6, 431.Notably,many other metal catalysts competitively isomerize 41 to Δ1,9-octalin, which is hydrogenated to trans-decalin 42b. However, this thermodynamiccontrol is not general. [Google Scholar]
  • Nishimura S.; Mochizuki F.; Kobayakawa S.Bull. Chem. Soc. Jpn.1970, 43, 1919. [Google Scholar]
  • Hatzellis K.; Pagona G.; Spyros A.; Demetzos C.; Katerinopoulos H. E.J. Nat. Prod.2004, 67, 1996. [PubMed] [Google Scholar]
  • González-Sierra M.; Laborde M. A.; Rúveda E. A.Synth. Commun.1987, 17, 431. [Google Scholar]
  • a. Wang L.-C.; Jang H.-Y.; Roh Y.; Lynch V.; Schultz A. J.; Wang X.; Krische M. J.J. Am. Chem. Soc.2002, 124, 9448. [PubMed] [Google Scholar]
    b. Hartung J.; Pulling M. E.; Smith D. M.; Yang D. X.; Norton J. R.Tetrahedron2008, 64, 11822. [Google Scholar]
    c. Leggans E. K.; Barker T. J.; Duncan K. K.; Boger D. L.Org. Lett.2012, 14, 1428. [PMC free article] [PubMed] [Google Scholar]
  • Lackner G. L.; Quasdorf K. W.; Overman L. E.J. Am. Chem. Soc.2013, 135, 15342. [PMC free article] [PubMed] [Google Scholar]
  • Mundal D. A.; Avetta C. T. Jr.; Thomson R. J.Nature Chem2010, 2, 294. [PubMed] [Google Scholar]
  • Syntheses of complex halogenated molecules:
    a. King S. M.; Calandra N. A.; Herzon S. B.Angew. Chem., Int. Ed.2013, 52, 3642. [PubMed] [Google Scholar]
    b. Nilewski C.; Carreira E. M.Eur. J. Org. Chem.2012, 1685. [Google Scholar]
    c. Chung W.-J.; Vanderwal C. D.. Acc. Chem.Res.; in press. [PMC free article] [PubMed]
    d. Snyder S. A.; Brucks A. P.; Treitler D. S.; Moga I.J. Am. Chem. Soc.2012, 134, 17714. [PubMed] [Google Scholar]
    Recent methods to synthesize chiral nonracemic organohalides:
    e. Hu D. X.; Shibuya G. M.; Burns N. Z.J. Am. Chem. Soc.2013, 135, 12960. [PubMed] [Google Scholar]
    f. Nicolaou K. C.; Simmons N. L.; Ying Y.; Heretsch P. M.; Chen J. S.J. Am. Chem. Soc.2011, 133, 8134. [PMC free article] [PubMed] [Google Scholar]

Articles from Journal of the American Chemical Society are provided here courtesy of American Chemical Society

Simple,
Chemoselective Hydrogenation with Thermodynamic
Stereocontrol (2024)

FAQs

What is the stereochemistry of hydrogenation? ›

The stereochemistry of the hydrogenation of cyclo- hexanones is governed by the von Auwers-Skita hydrogenation rule' which states that the hydro- genation over Pt catalyst in acid medium produces predominantly the cis alcohols, while the truns alcohols are favoured in neutral or basic medium.

What is the thermodynamic of co2 hydrogenation to methanol? ›

Reaction thermodynamics. Hydrogenation of carbon dioxide to methanol is an exothermic reaction with experimentally determined enthalpy change ΔH298 K = −48.98 kJ/mol [37]. As two molecules of products form from four molecules of reactants, low temperature and high pressure favour the production of methanol.

Is hydrogenation syn or anti? ›

Hydrogenation of alkenes is a syn addition.

Is hydrogenation Regioselective? ›

Addition of hydrogen to a carbon-carbon double bond is called hydrogenation. The overall effect of such an addition is the reductive removal of the double bond functional group. Regioselectivity is not an issue, since the same group (a hydrogen atom) is bonded to each of the double bond carbons.

What are the two types of stereochemistry? ›

Stereoisomers can be broadly classified into two types, namely enantiomers and diastereomers.
  • Enantiomers. When two isomers are mirror images of each other, the type of isomerism is called enantiomerism and these isomers are referred to as enantiomers. ...
  • Diastereomers.

What are the two types of hydrogenation? ›

Hydrogenated catalysts generally fall into two broad categories: hom*ogeneous catalysts and heterogeneous catalysts. The hom*ogeneous catalyst can be dissolved in a solvent containing an unsaturated substrate.

What catalyst is used to convert CO2 to methanol? ›

Catalysts with different metals like Cu, Zn, Ag, Cr, and Pd have been employed for CO2 hydrogenation to methanol (Kattel et al., 2017a; Dang et al., 2019b; Din et al., 2019). Nevertheless, Cu-based catalysts exhibit high activity and selectivity.

What does CO2 hydrogenation to methanol yield? ›

A yield of 25 % per pass is obtained. The permeate and the water- methanol mixture from the phase separator is finally separated in a distillation column.

What catalyst is used in the hydrogenation of carbon monoxide to methanol? ›

Alcohol-Assisted Hydrogenation of Carbon Monoxide to Methanol Using Molecular Manganese Catalysts | JACS Au.

What is hydrogenation and what is bad about it? ›

Hydrogenated vegetable oils are widely used in the food industry to improve the taste and texture of processed foods. Still, they harbor trans fats, which may negatively affect heart health, inflammation, and blood sugar control.

What is a simple example of hydrogenation? ›

Hydrogenation is generally carried out at lower temperatures. It is used to preserve raw materials or ingredients such as Alcohol, Ammonia, various polymers, etc. Hydrogenation is a process used to convert vegetable oil from a liquid to a solid or semi-solid fat.

What is the reverse of hydrogenation? ›

In chemistry, dehydrogenation is a chemical reaction that involves the removal of hydrogen, usually from an organic molecule. It is the reverse of hydrogenation.

What is chemoselective regioselective Stereoselective? ›

Regioselectivity is when the two possible products in the reaction are regioisomers (also called constitutional isomers) Stereoselectivity is when the two possible products in the reaction are stereoisomers. Chemoselectivity is when the reactants will prefer one functional group over another in the substrate.

Why is hydrogenation stereospecific? ›

A catalyst (Pd/C, PtO2, Ni0) is needed to weaken the σH−H bond, and the reaction is normally heterogeneous (occurs on a surface). The reaction is stereospecific since hydrogen is added in a concerted fashion from the same face of the alkene, without any opportunity for bond rotation.

Does hydrogenation have stereochemistry? ›

Hydrogenation usually occurs with syn stereochemistry: both hydrogens add to the double bond from the same face.

What is the stereochemistry of hydrohalogenation? ›

Hydrohalogenation results in the Markovnikov addition of a hydrogen (less substituted side) and a halogen (Cl, Br, or I--more substituted side) across an alkene forming an alkyl halide. There is no stereospecificity associated with this reaction, but the intermediate is a carbocation so rearrangements are possible.

Is hydrogenation stereospecific or stereoselective? ›

Note 5. The catalytic hydrogenation of alkenes with Pd/C and H2 is another example of a reaction that is generally stereoselective (for syn addition to alkenes) but in practice is not stereospecific.

What is the structure of hydrogenation? ›

Hydrogenation is the process where hydrogen atoms bind to the double bond of a compound, facilitating its conversion to a single bond, in the presence of a catalyst. Hydrocarbons with double bonds are classified as unsaturated. Those that only contain single bonds are saturated.

Does hydrogenation form enantiomers? ›

Hydrogenation processes are interesting by themselves, but as the two hydrogen atoms are not added to the same atom of the unsaturated bond but one to one, depending on the substrate substitution it is possible to achieve the preferential formation of one enantiomer over the other, leading to mixtures with high ...

References

Top Articles
Latest Posts
Article information

Author: Laurine Ryan

Last Updated:

Views: 5419

Rating: 4.7 / 5 (57 voted)

Reviews: 80% of readers found this page helpful

Author information

Name: Laurine Ryan

Birthday: 1994-12-23

Address: Suite 751 871 Lissette Throughway, West Kittie, NH 41603

Phone: +2366831109631

Job: Sales Producer

Hobby: Creative writing, Motor sports, Do it yourself, Skateboarding, Coffee roasting, Calligraphy, Stand-up comedy

Introduction: My name is Laurine Ryan, I am a adorable, fair, graceful, spotless, gorgeous, homely, cooperative person who loves writing and wants to share my knowledge and understanding with you.