a Molar ratio of substrate to catalyst.b 1H NMR yield with 1,3,5-trimethoxybenzene as
an internal standard. c Determined by1H NMR.d HCO2H (4 equiv).e HCO2H (12 equiv).
A number of cis -2-acyl cycloalkane-1-carboxylic acids were
synthesized via either
AlCl3-catalyzed
Friedel-Crafts acylation19 or Ni-catalyzed
cross-coupling of anhydrides with organozinc
reagents,20 followed by submitting them to the optimal
conditions (Table 2). The stereochemistry of products was further
demonstrated by the one-dimensional NOE (Nuclear Overhauser Effects)
spectra of four randomly selected representative compounds (2c ,2x , 2aa , and 2b ). We focused our substrate
scope on 2-aroyl substrates (R = aryl, cis -1 ), to which
previously established syn -selective reductive lactonization
protocols were not applicable.10b In addition tocis -1a , other 2-aroyl substituted substrates were
converted to the desired diastereopure products in 60-99% yields and
98:2 to >99:1 diastereoselectivity, strongly favoring the
formation of syn -isomers (2a -2u ). The
substituents, for instance, alkyls, phenyl, phenyloxyl, benzenesulfenyl,
and halogen atoms, on phenyl rings, whether electron-donating or
withdrawing, all uniformly gave ≥ 98:2 dr (2a -2n ),
although the yields varied. Substrates bearing more complex arene rings
such as 5,6,7,8-tetrahydronaphthalen-2-yl (2o ),
4-methylnaphthalen-2-yl (2p ), naphthalen-2-yl and
naphthalen-1-yl (2q ), pyren-4-yl (2r ),
9H -fluoren-3-yl (2s ), and
dibenzo[b ,d ]thiophen-2-yl (2u ) also
transformed in >99:1 dr. The ketoacid with
dibenzo[b ,d ]furan-2-yl (2t ) was reductively
lactonized in ≥ 98:2 dr values. Oxidation of the sulfur atom ofsyn -2u did not reduce its diastereopurity.
We next examined the effects of backbone architectures on the
stereochemial results. Introduction of an endocyclic disubtituted
(2w ) or tetrasubstituted alkene moiety (2x ), being
fused with a bridged bicyclic [2.2.1] or [2.2.2] scaffold
(2y and 2z ), or contraction of the cyclohexane ring to
cyclopentane (2aa ) or cyclobutane (2ab ), did not
affect the stereochemical outcomes. All products were produced in
>99:1 dr values, and syn -products were formed
exclusively.
Substrates with alkanoyls (R = alkyl, cis -1 ) were also
amenable to the diastereocontrol. The reactions ofcis -1b and cis -1ac both gave
> 99:1 dr values and reasonable yields.
Table 2 Substrate scopea
a cis -2a , 2c –2h ,2k –2o , 2q, 2w, 2ab were isolated by
extraction. b The regioselective ratio (rr) of
ketoacid was 94:6 (para :ortho ), and it was inherited to
the next step. c The rr of ketoacid and
product was 50:50. d 0.1 mmol of ketoacid was
used. e HCO2H (16 equiv.),C1 (S/C = 500), H2O/EtOH (1:1), 80oC, 6 h. f The rr of
ketoacid and product was 96:4.g HCO2H (12 equiv.), C1(S/C = 500), H2O/EtOH (1:1), 80oC, 6 h.
Mechanistic studies were performed
(Scheme 2). The isotope tracing experiments with HCO2D,
DCO2D, and D2O (Scheme 2a) demonstrated
that (1) the transfered hydride originated from the formyl group of
formic acid, and (2) that H–D exchange, both between iridium hydride
([Ir]–H) and D2O and between iridium deteride
([Ir]–D) and H2O, occurred before the hydride or
deuteride transfer.12, 15a, 21 The H–D exchange
between [Ir]–D and solvent was dramatically suppressed by using
DCO2D and D2O-EtOH as the respective
deteride source and solvent, and thereby a 95% D-incorporation was
obtained. We also measured the kinetic isotope effect (KIE), and akH /kD = 4.8 was observed
(Scheme 2b). The primary KIE implied that the rate-determining step
should involve a C–H cleavage event.
The key intermediate during the reductive lactonization was observed by1H NMR of the crude reaction mixture (Scheme 2c). It
turned out to be an alcohol (3a ), which was generated from the
diastereospecific reduction of the ketone moiety. Treatment of this
alcohol under acidic conditions (AcOH) yielded lactonesyn -2a with >99:1 syn :antiratio. We also found that the reduction of 4-ketoestercis -1m’ did not occur (Scheme 2d), which indicated the
important role of the carboxylic acid group in facilitating the
reduction. Similar carboxylic acid effect was also observed
by Rovis and
coworkers.10b
Scheme 2 Isotope labelling, KIE, and other mechanistic
expecriments.
Scheme 3 Proposed mechanism and diastereocontrol models
On the basis of the mechanistic studies, a plausible mechanism was
proposed (Scheme 3a). In light of the KIE studies, the generation of
iridium hydride B via β-hydride elimination was suggested to be
the rate-determining step.22 Subsequent hydride
transfer to the carbonyl group, which was activated by the carboxylic
group, served as the diastereo-determining step. Considering the
experimental results in Scheme 2d, we proposed that the hydrogen bonding
between the carboxylic acid and ketone moiety accounted for the
activation. In detail, protonation of the C=O bond of carboxylic acid1a increased its acidity, and an intramolecular hydrogen
bonding between O–H and carbonyl C=O occurred, as designated by the
structure C . Although C could resonate withD , it was suggested as the major contributor for explaining the
carboxylic effect. On the other hand, the intramolecular hydrogen
bonding, to some extent, fixed the configuration by connecting the
ketone and carboxylic acid moieties. The fixed and somewhat rigid
configuration of the molecule rendered the hydride transfer more
stereoselective. In this regard, it is the intramolecular hydrogen
bonding that holds responsibility for the excellent diastereocontrol in
the hydride transfer step. Once diastereospecifically formed,3a immediately undergoes intramolecular esterification,
following classic acid-catalyzed addition-elimination mechanism, to form
lactone syn -2a , without eroding the stereochemistry of
the newly generated stereocenter.
To better understand the diastereocontrol of the hydride transfer, we
performed the DFT calculations. The calculated fixed configurations and
Bürgi-Dunitz angles of attack for hydride delivery to carbonyl were
shown in Scheme 3b and 3c.23 In the case that the
carboxylic assumes axial position and benzoyl equatorial position
(Scheme 3b), the frontside attack will encounter severe steric repulsion
against two axial hydrogen atoms and one axial carboxylic group.
However, the backside attack faces very little steric retardation.
Consequently, the hydride delivery in this direction kinetically
preferentially gave syn -2a . Similar stereocontrol is
also applicable in the other configuration bearing an equatorial
carboxylic and an axial benzoyl (Scheme 3c). In one word, it is the
steric repulsion between sterically bulky iridium hydride and the
cycloalkyl ring that governs the diastereoselectivity.
To explain the backbone architecture effects in Table 2 and also to
predict the stereochemical outcomes of the potential reactions of
substrates with similar stereochemistry, we proposed a rule of thumb
(Scheme 3d). The hypothesis that the hydrogen bond form an additional
ring (ring B) leads to a cis -fused bicyclic system, and thereby
the hydride reduction of the “pseudo-endocyclic” carbonyl was endowed
with facial selectivity—the hydride can attack from either convex or
concave face.24 Owing to the large steric hindrance of
the iridium hydride, it would preferentially donate its hydride from the
sterically much more accessible convex face, in this way generating a
configuration with all the hydrogen atoms on the three tertiary
stereocenters residing in the same convex face.
Scheme 4 Gram-scale reactions and synthetic applications
In the two gram-scale reactions, 2a and 2m were both
isolated in excellent yields and > 99:1 dr (Scheme 4a). The
purification was very convenient. Column chromatography was not
required. Extraction with ethyl acetate followed by drying and
concentration afforded desired products in good NMR purities (see
Supporting Information). The 4-bromophenyl group in 2m acted a
handle for further chemical manipulations (Scheme 4b). For examples, the
Suzuki coupling of 2m with three aryl boric acids carrying
electron-withdrawing or electron-donating substituents delivered more
functionalized products 4a -4c in 56-96% yields, and
the stereochemistry of the three stereocenters was not affected.
trans -2-Benzoylcyclohexane-1-carboxylic acid
(trans -1a ) did not undergo the reductive lactonization
(Scheme 4c), highlighting the importance of the configurational effect
of the substrate. In other words, our iridium catalysts showed excellent
capability in discriminating the cis - andtrans -substrates, and highly selectively catalyzed the reactions
of cis -substrates, leaving the trans -ones intact. This
selectivity provides an easy procedure to separate the cis - andtrans -2-acylcycloalkane-1-carboxylic acids, which were generated
as a diasteremeric mixture in some cases and were difficult to be
separated due to their almost identical polarity. As shown in Scheme 4d,
subjection of equimolar cis -1a andtrans -1a to standard reductive latonization conditions
produced cis -2b in 94% yield andtrans -1a in 92% recovery. Due to the largely different
polarities and water solubility of cis -2b andtrans -1a , they were easily separated.
Conclusions
We have developed a highly diastereoselective method for efficient
synthesis of diastereopure bi- and polycyclic γ-lactones, usingcis -2,3-fused 4-oxo-butanoic acids as the starting materials.
This method features the use of a
[Cp*IrIIICl/PyIm]+Cl-catalyst with formic acid as the hydride source. Advantages of this
method include excellent diastereoselectivity control, use of
water-ethanol as solvent, broad substrate scope, and high catalyst
efficiency (S/C up to 5000). The gram-scale reactions take place
efficiently in excellent yields and diastereocontrol. Mechanistic
studies suggested that the iridium hydride formation be the
rate-determining step, and that the hydride transfer step be the
diastereo-determining step. The large steric hindrance of the iridium
hydride species underlies the success of diastereocontrol. The
carboxylic acid group of substrates plays important roles in activating
the substrates and in rendering a relatively rigid configuration to
highly diastereoselectively receive the hydride. DFT calculations
provide detailed insights into the nature of diastereocontrol, from the
perspectives of configurational analysis and Duniz angles of attack. An
empirical rule based on facial selectivity analysis for explaining and
predicting the stereochemistry is also proposed. Our iridium catalysts
only work on cis -2,3-fused 4-oxo-butanoic acids, showing
excellent level of molecular recognization. This selectivity can be
harnessed to separate the diastereomeric mixtures of cis - andtrans -2-acylcycloalkane-1-carboxylic acids. Compared with
previous methods, our method shows superiority in terms of substrate
scope, degree of diastereocontrol, and sustainability.
Experimental
To a 10-mL reaction tube was
sequentially added cis -2-acyl cycloalkane-1-carboxylic acids1 (0.2 mmol), ethanol (0.5 mL), 1 mL of C1 solution in
deionized water (0.0001 mol/L, S/C = 2000). The tube was then
sealed with a rubber cap. A syringe needle was inserted into the cap to
connect the inner to outer atmosphere. The mixture was stirred for 3
minutes in an 80 °C heating block, followed by addition of formic acid
(60 μL, 1.6 mmol, 8 equiv) in one portion via a microsyringe. After
stirring for 2 h, the reaction mixture was cooled to room temperature,
diluted with saturated brine (2 mL), and extracted with ethyl acetate (2
mL × 3). The organic phase was dried with anhydrous
Na2SO4 and then removed under reduced
pressure. 1H NMR pure syn -2a ,2c –2h , 2k –2o , 2q ,2w , 2ab were obtained without further purification.
Other products were purified by silica gel column chromatography.
Supporting Information
The supporting information for this article is available on the WWW
under https://doi.org/10.1002/cjoc.2023xxxxx.
Acknowledgement (optional)
This research was supported by the National Key Research and Development
Program of China (No.2022YFF0709803), Natural Science Foundation of
Beijing Municipality (no. 2202041), and the High Performance Computing
Platform of Beijing University of Chemical Technology (BUCT).
References
- Jiwajinda, S.; Santisopasri, V.; Murakami, A.; Kawanaka, M.; Kawanaka,
H.; Gasquet, M.; Eilas, R.; Balansard, G.; Ohigashi, H. In vitro
anti-tumor promoting and anti -parasitic activities of the
quassinoids from Eurycoma longifolia, a medicinal plant in Southeast
Asia. J. Ethnopharmacol . 2002 , 82 , 55-58.
- Huang, J.-m.; Yokoyama, R.; Yang, C.-s.; Fukuyama, Y. Merrilactone A,
a novel neurotrophic sesquiterpene dilactone from Illicium
merrillianum. Tetrahedron Lett . 2000 , 41 ,
6111-6114.
- (a) Duh, C.-Y.; Wang, S.-K.; Chia, M.-C.; Chiang, M. Y. A novel
cytotoxic norditerpenoid from the Formosan soft coral Sinularia
inelegans. Tetrahedron Lett. 1999 , 40 ,
6033-6035; (b) Cheng, S.-Y.; Chuang, C.-T.; Wen, Z.-H.; Wang, S.-K.;
Chiou, S.-F.; Hsu, C.-H.; Dai, C.-F.; Duh, C.-Y. Bioactive
norditerpenoids from the soft coral Sinularia gyrosa. Bioorg.
Med. Chem . 2010 , 18 , 3379-3386.
- Yasuda, N.; Sugimoto, Y.; Kato, M.; Inanaga, S.; Yoneyama, K.
(+)-Strigol, a witchweed seed germination stimulant, from Menispermum
dauricum root culture. Phytochemistry 2003 , 62 ,
1115-1119.
- Zhang, X.; Meng, R.; Wang, H.; Xing, J. Differential Effects of
Components in Artemisia annua Extract on the Induction of
Drug-Metabolizing Enzyme Expression Mediated by Nuclear Receptors.Planta Med. 2020 , 86 , 867-875.
- Barnes, E. C.; Choomuenwai, V.; Andrews, K. T.; Quinn, R. J.; Davis,
R. A. Design and synthesis of screening libraries based on the
muurolane natural product scaffold. Org. Biomol. Chem.2012 , 10 , 4015-4023.
- Klepp, J.; Sadgrove, N. J.; Legendre, S. V. A. M.; Sumby, C. J.;
Greatrex, B. W. Biomimetic Synthesis of Mitchellenes B–H from the
Abundant Biological Precursor 14-Hydroxy-6,12-muuroloadien-15-oic
Acid. J. Org. Chem. 2019 , 84 , 9637-9647.
- For selected reviews, see: (a) Schall, A.; Reiser, O. Synthesis of
Biologically Active Guaianolides with a trans-Annulated Lactone
Moiety. Eur. J. Org. Chem . 2008 , 2008 ,
2353-2364; (b) Gil, S.; Parra, M.; Rodriguez, P.; Segura, J. Recent
Developments in γ-Lactone Synthesis. Mini-Rev. Org. Chem.2009 , 6 , 345-358; (c) Neumeyer, M.; Brückner, R.
Nonracemic γ-Lactones from the Sharpless Asymmetric Dihydroxylation of
β,γ-Unsaturated Carboxylic Esters. Eur. J. Org. Chem.2016 , 2016 , 5060-5087; (d) Mao, B.; Fañanás-Mastral,
M.; Feringa, B. L. Catalytic Asymmetric Synthesis of Butenolides and
Butyrolactones. Chem. Rev . 2017 , 117 ,
10502-10566; (e) Sartori, S. K.; Diaz, M. A. N.; Diaz-Muñoz, G.
Lactones: Classification, synthesis, biological activities, and
industrial applications. Tetrahedron 2021 , 84 ,
132001.
- For selected synthetic examples, see: (a) Chou, Y.-Y.; Peddinti, R.
K.; Liao, C.-C. Highly Diastereoselective and Asymmetric Diels−Alder
Reactions of Masked o-Benzoquinones with Chiral Racemic and Homochiral
Furans. Synthesis of Optically Active Tricyclic γ-Lactones. Org.
Lett. 2003 , 5 , 1637-1640; (b) Barros, M. T.; Burke,
A. J.; Lou, J.-D.; Maycock, C. D.; Wahnon, J. R. Highly
Stereoselective Aldol Reaction for the Synthesis of γ-Lactones
Starting from Tartaric Acid. J. Org. Chem . 2004 ,69 , 7847-7850; (c) Powell, L. H.; Docherty, P. H.; Hulcoop, D.
G.; Kemmitt, P. D.; Burton, J. W. Oxidative radical cyclisations for
the synthesis of γ-lactones. Chem. Commun. 2008 ,
2559-2561; (d) Kalmode, H. P.; Handore, K. L.; Reddy, D. S. Access to
Fused Tricyclic γ-Butyrolactones, A Natural Product-like Scaffold.J. Org. Chem . 2017 , 82 , 7614-7620; (e) Truax,
N. J.; Ayinde, S.; Van, K.; Liu, J. O.; Romo, D.
Pharmacophore-Directed Retrosynthesis Applied to Rameswaralide:
Synthesis and Bioactivity of Sinularia Natural Product Tricyclic
Cores. Org. Lett. 2019 , 21 , 7394-7399; (f)
Deng, M.; Zhang, X.; Li, Z.; Chen, H.; Zang, S.; Liang, G. Rapid
Construction of the Common [5–5–6] Tricyclic Ring Skeleton in
Polycyclic Cembranoids and Norcembranoids via Intramolecular
1,3-Dipolar Cycloaddition. Org. Lett . 2019 , 21 ,
1493-1496.
- (a) Pourahmady, N.; Eisenbraun, E. J. (±)-3-Phenylhexahydrophthalides.
Synthesis and structural assignments. J. Org. Chem.1983 , 48 , 3067-3070; (b) Bercot, E. A.; Kindrachuk, D.
E.; Rovis, T. Complementary Diastereoselective Reduction of Cyclic
γ-Keto Acids: Efficient Access to Trisubsituted γ-Lactones.Org. Lett. 2005 , 7 , 107-110; (c) Eisenbraun, E.
J.; Browne, C. E.; Irvin-Willis, R. L.; McGurk, D. J.; Eliel, E. L.;
Harris, D. L. Structure and stereochemistry 4aβ,7α,7aβ-Nepetalactone
from Nepeta mussini and its relationship to the 4aα,7α,7aα- and
4aα,7α,7aβ-Nepetalactones from N. cataria. J. Org. Chem.1980 , 45 , 3811-3814; (d) Seiji, M.; Nobuhiro, A.;
Fumiko, F.; Kunihiro, S.
3-Phenylhexahydro-1(3H )-isobenzofuranones. Heterocycles1987 , 26 , 1813-1826.
- Yang, Z.; Zhu, Z.; Luo, R.; Qiu, X.; Liu, J.-t.; Yang, J.-K.; Tang, W.
Iridium-catalyzed highly efficient chemoselective reduction of
aldehydes in water using formic acid as the hydrogen source.Green Chem . 2017 , 19 , 3296-3301.
- Liu, J.-t.; Yang, S.; Tang, W.; Yang, Z.; Xu, J. Iridium-catalyzed
efficient reduction of ketones in water with formic acid as a hydride
donor at low catalyst loading. Green Chem. 2018 ,20 , 2118-2124.
- (a) Xu, D.; Chen, Y.; Liu, C.; Xu, J.; Yang, Z. Iridium-catalyzed
highly chemoselective and efficient reduction of nitroalkenes to
nitroalkanes in water. Green Chem . 2021 , 23 ,
6050-6058; (b) Wang, T.; Liu, C.; Xu, D.; Xu, J.; Yang, Z.
Iridium-Catalyzed and pH-Dependent Reductions of Nitroalkenes to
Ketones. Molecules 2022 , 27 , 7822.
- Liu, C.; Chen, Y.; Yang, Z. Iridium-Catalyzed Stereoselective Transfer
Hydrogenation of 1,5-Benzodiazepines. J. Org. Chem.2022 , 87 , 12001-12018.
- (a) Yang, S.; Tang, W.; Yang, Z.; Xu, J. Iridium-Catalyzed Highly
Efficient and Site-Selective Deoxygenation of Alcohols. ACS
Catal. 2018 , 8 , 9320-9326; (b) Wang, T.; Chen, Y.;
Chen, N.; Xu, J.; Yang, Z. Iridium-catalyzed highly stereoselective
deoxygenation of tertiary cycloalkanols: stereoelectronic insights and
synthetic applications. Org. Biomol. Chem. 2021 ,19 , 9004-9011; (c) Wang, J.; Wang, T.; Du, H.; Chen, N.; Xu,
J.; Yang, Z. Accessing para-Alkylphenols via Iridium-Catalyzed
Site-Specific Deoxygenation of Alcohols. J. Org. Chem.2023 , 88 , 12572-12584.
- (a) Li, J.; Tang, W.; Ren, D.; Xu, J.; Yang, Z. Iridium-catalysed
highly selective reduction–elimination of steroidal 4-en-3-ones to
3,5-dienes in water. Green Chem . 2019 , 21 ,
2088-2094; (b) Wang, T.; Miao, R.; Luo, R.; Xu, J.; Yang, Z.
Furan-2-yl Anions as γ-Oxo/Hydroxyl Acyl Anion Equivalents Enabled by
Iridium-Catalyzed Chemoselective Reduction. Org. Lett.2023 , 25 , 4705-4710. (c) Chen, Y.; Zhang, J.; Du, H.;
Luo, R.; Xu, J.; Yang, Z. Iridium-catalyzed reductive γ-lactonization
of ortho-acylbenzoic acids in water: sustainable access to phthalides.Org. Chem. Front . 2024 , DOI: 10.1039/D3QO02019C.
- For a review, see: (a) Wei, Y.; Liang, Y.; Luo, R.; Ouyang, L. Recent
advances of Cp*Ir complexes for transfer hydrogenation: focus on
formic acid/formate as hydrogen donors. Org. Biomol. Chem .2023 , 21 , 7484-7497; For selected work, see: (b) Xia,
Y.; Wang, S.; Miao, R.; Liao, J.; Ouyang, L.; Luo, R. Synthesis of
N-alkoxy amines and hydroxylamines via the iridium-catalyzed transfer
hydrogenation of oximes. Org. Biomol. Chem. 2022 ,20 , 6394-6399; (c) Ouyang, L.; Xia, Y.; Liao, J.; Luo, R.
One-Pot Transfer Hydrogenation Reductive Amination of Aldehydes and
Ketones by Iridium Complexes “on Water”. Eur. J. Org. Chem.2020 , 2020 , 6387-6391; (d) Liu, L.; Luo, R.; Tong, J.;
Liao, J. Iridium-catalysed reductive allylic amination of
α,β-unsaturated aldehydes. Org. Biomol. Chem . 2024 ,22 , 585-589; (e) Ouyang, L.; Miao, R.; Yang, Z.; Luo, R.
Iridium-catalyzed reductive amination of carboxylic acids. J.
Catal. 2023 , 418 , 283-289; (f) Ouyang, L.; Liang, Y.;
Wang, S.; Miao, R.; Liao, J.; Yang, Z.; Luo, R. Iridium-catalyzed
reductive hydroamination of terminal alkynes in water. J.
Catal. 2023 , 427 , 115096.
- (a) Song, H.-Y.; Jiang, J.; Wu, C.; Hou, J.-C.; Lu, Y.-H.; Wang,
K.-L.; Yang, T.-B.; He, W.-M. Semi-heterogeneous
g-C3N4/NaI dual catalytic C–C bond
formation under visible light. Green Chem . 2023 ,25 , 3292-3296; (b) Lu, Y.-H.; Mu, S.-Y.; Li, H.-X.; Jiang, J.;
Wu, C.; Zhou, M.-H.; Ouyang, W.-T.; He, W.-M. EtOH-catalyzed
electrosynthesis of imidazolidine-fused sulfamidates from N-sulfonyl
ketimines, N-arylglycines and formaldehyde. Green Chem .2023 , 25 , 5539-5542; (c) Ouyang, W.-T.; Ji, H.-T.;
Jiang, J.; Wu, C.; Hou, J.-C.; Zhou, M.-H.; Lu, Y.-H.; Ou, L.-J.; He,
W.-M. Ferrocene/air double-mediated
FeTiO3-photocatalyzed semi-heterogeneous annulation of
quinoxalin-2(1H )-ones in EtOH/H2O. Chem.
Commun . 2023 , 59 , 14029-14032.
- Dietliker, K.; Murer, P.; Huesler, R.; Jung, T. Absorbing light 320 -
450 nm α-hydroxyketones as photoinitiators for clear coats.WO2008122504 , 2008 .
- Bercot, E. A.; Rovis, T. Highly Efficient Nickel-Catalyzed
Cross-Coupling of Succinic and Glutaric Anhydrides with Organozinc
Reagents. J. Am. Chem. Soc. 2005 , 127 , 247-254.
- (a) Wang, C.; Chen, H.-Y. T.; Bacsa, J.; Catlow, C. R. A.; Xiao, J.
Synthesis and X-ray structures of cyclometalated iridium complexes
including the hydrides. Dalton Trans. 2013 , 42 ,
935-940; (b) Onishi, N.; Ertem, M. Z.; Xu, S.; Tsurusaki, A.; Manaka,
Y.; Muckerman, J. T.; Fujita, E.; Himeda, Y. Direction to practical
production of hydrogen by formic acid dehydrogenation with Cp*Ir
complexes bearing imidazoline ligands. Catal. Sci. Technol.2016 , 6 , 988-992.
- (a) Gómez-Gallego, M.; Sierra, M. A. Kinetic Isotope Effects in the
Study of Organometallic Reaction Mechanisms. Chem. Rev.2011, 111 , 4857-4963; (b) Janak, K. E., in
Comprehensive Organometallic Chemistry III, eds. D. M. P. Mingos and
R. H. Crabtree, Elsevier, Oxford, 2007 , pp. 541-571; (c)
Truong, P. T.; Miller, S. G.; McLaughlin Sta. Maria, E. J.; Bowring,
M. A. Large Isotope Effects in Organometallic Chemistry. Chem.
Eur. J. 2021 , 27 , 14800-14815.
- For selected reviews: (a) Ashby, E. C.; Laemmle, J. T. Stereochemistry
of organometallic compound addition to ketones. Chem. Rev.1975 , 75 , 521-546; (b) Gung, B. W. Structure
Distortions in Heteroatom-Substituted Cyclohexanones, Adamantanones,
and Adamantanes: Origin of Diastereofacial Selectivity. Chem.
Rev . 1999 , 99 , 1377-1386; (c) Ohwada, T.
Orbital-Controlled Stereoselections in Sterically Unbiased Cyclic
Systems. Chem. Rev. 1999 , 99 , 1337-1376; (d)
Wipf, P.; Jung, J.-K. Nucleophilic Additions to 4,4-Disubstituted
2,5-Cyclohexadienones: Can Dipole Effects Control Facial Selectivity?Chem. Rev . 1999 , 99 , 1469-1480; For selected
work, see: (e) Bürgi, H. B.; Dunitz, J. D.; Lehn, J. M.; Wipff, G.
Stereochemistry of reaction paths at carbonyl centres.Tetrahedron . 1974 , 30 , 1563-1572; (f) Bürgi, H.
B.; Dunitz, J. D.; Shefter, E. Geometrical reaction coordinates. II.
Nucleophilic addition to a carbonyl group. J. Am. Chem. Soc.1973 , 95 , 5065-5067.
- For reviews, see: (a) Kaselj, M.; Chung, W.-S.; le Noble, W. J. Face
Selection in Addition and Elimination in Sterically Unbiased Systems.Chem. Rev. 1999 , 99 , 1387-1414; (b) Mehta, G.;
Chandrasekhar, J. Electronic Control of Facial Selection in Additions
to Sterically Unbiased Ketones and Olefins. Chem. Rev .1999 , 99 , 1437-1468; For selected examples, see: (c)
Dols, P. P. M. A.; Arnouts, E. G.; Rohaan, J.; Klunder, A. J. H.;
Zwanenburg, B. Nucleophilic additions to tricyclodecadienone epoxides.
The Payne rearrangement of α,β-epoxycyclopentanols contained in a
rigid tricyclic system. Tetrahedron . 1994 , 50 ,
3473-3490; (d) Dhakal, R.; Ivancic, M.; Brewer, M. Two-Step Sequence
of Cycloadditions Gives Structurally Complex Tetracyclic
1,2,3,4-Tetrahydrocinnoline Products. J. Org. Chem.2018 , 83 , 6202-6209.