You are on page 1of 5

Note

Cite This: J. Org. Chem. 2018, 83, 10688−10692 pubs.acs.org/joc

Mechanism of SmI2 Reduction of 5‑Bromo-6-oxo-6-phenylhexyl


Methanesulfonate Studied by Spin Trapping with 2‑Methyl-2-
nitrosopropane
Christopher D. Aretz, Joseph E. McPeak, Gareth R. Eaton, Sandra S. Eaton, and Bryan J. Cowen*
Department of Chemistry and Biochemistry, University of Denver, Denver, Colorado 80208, United States
*
S Supporting Information

ABSTRACT: The radical formed by reduction of 5-bromo-6-oxo-6-


See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

phenylhexyl methanesulfonate, an α-bromoketone, with SmI2 was spin


trapped with 2-methyl-2-nitrosopropane. Electron paramagnetic resonance
spectra of the spin adduct and the adduct formed in the analogous reaction
with selectively deuterated substrate identify the radical intermediate in this
Downloaded via UNIV DE CHILE on May 17, 2019 at 19:57:06 (UTC).

SmI2 reduction as a carbon-centered radical. This result supports the proposal that the formation of reactive Sm-enolates arises
from reduction of the carbon−bromine bond rather than a ketyl radical anion.

S amarium(II) iodide (SmI2) is a powerful single electron


reducing agent that can be used under mild conditions to
perform a wide range of synthetically useful reactions.1 Because
Based on previous work pioneered by Molander and
coworkers, two scenarios are proposed that could lead to
samarium enolate 5 (Scheme 2).6 In pathway A, donation of a
its utility in organic synthesis was first shown by Kagan in
1977, particular attention has focused on reductive carbon− Scheme 2. Mechanistic Scenarios for Enolate Generation
carbon bond forming processes promoted by the reagent.
Samarium enolate generation through reduction of an α-
halogenated or α-oxygenated carbonyl compound has been
shown to be an efficient method for Reformatsky aldol-type
reactions.2 The mechanism of reduction likely depends on
various factors, including the type of carbonyl compound, type
of halogen, nature of samarium reagent, and additives.3
Additionally, chelation of the carbonyl oxygen and α-
heteroatom with the samarium center could influence the
reduction as previously established.4 We recently reported that
enolates generated from action of SmI2 can be exploited in an
intramolecular alkylation reaction of a primary mesylate such single electron from Sm(II) to substrate 1 leads to ketyl radical
as type 1 and the analogous intramolecular Reformatsky aldol anion 6. This species can be further reduced by an additional
cyclization of aldehyde type 2 (Scheme 1).5 We now wish to equivalent of Sm(II) to dianionic intermediate 7 that leads to
probe the mechanism of enolate generation from α-bromo aryl elimination of the α-bromo group and formation of the enolate
ketones by SmI2, which were utilized for the synthesis of 5. In pathway B, single electron reduction of the α-bromo
cyclopentane products of type 3 and 4. group directly would lead to a radical anion that upon
fragmentation of the C−Br bond would give α-centered radical
Scheme 1. Reductive Cyclization Reactions Promoted by 8. This intermediate would be reduced further to the enolate 5
SmI2 by Sm(II). Although this question was raised in 1986, it
remains unanswered despite the widespread use of SmI2. We
hypothesized that formation of ketyl radical 6 could be
distinguished from α-centered radical 8 by spin trapping with
MNP (2-methyl-2-nitrosopropane).7 The trapped products
would display distinctive hyperfine coupling patterns in their
EPR spectra, leading to a diagnostic identification of the parent
radical species generated from initial reaction of substrate 1
with SmI2.

Received: June 15, 2018


Published: August 13, 2018

© 2018 American Chemical Society 10688 DOI: 10.1021/acs.joc.8b01517


J. Org. Chem. 2018, 83, 10688−10692
The Journal of Organic Chemistry Note

We initiated our study by analyzing the reaction between α- The much-better resolved spectrum for the deoxygenated
bromoketone substrate 9a and SmI2 in the presence of sample is shown in Figure 2B. The value of the nitrogen
trapping agent MNP by EPR spectroscopy (Scheme 3). The

Scheme 3. Intermediate Radical Trapping with MNP

reaction of an unstable radical with a spin trap rapidly forms a


more stable radical that can be studied by EPR.8 In fact, the
bimolecular reaction rate of an alkyl radical with MNP has
been measured to be on the order of 108 M−1 s−1.9 For
comparison, the bimolecular reaction rate of an alkyl radical
with SmI2−hexamethylphosphoramide (HMPA) has been
measured to be ∼106 M−1 s−1 and would be significantly
slower without the HMPA additive.10 The spectrum of the
species that was trapped by MNP during the reaction of 9a
with SmI2 is shown in Figure 1B. The three-line pattern is due
to hyperfine coupling to the nitrogen that is present in all
radicals trapped with MNP.11,12 Although spin trapping
experiments typically are performed in air-saturated solution,
the trapped adduct in this case is sufficiently stable that the
solution could be degassed by freeze−pump−thaw methods.

Figure 2. X-band EPR spectra of spin-trapped adducts in


deoxygenated THF after workup to remove SmI2. Spectra were
collected with 0.54 μT B1 and 110 s acquisition time. (A) Adduct 12
from reaction of 11. (B) Adduct 10a from reaction of 9a. (C) Adduct
10b from reaction of 9b. NMR spectra of 9b found a 16:1 ratio of
deuteron to proton at the α-carbon. Simulation of the EPR spectrum
in part C is consistent with this ratio of 10b:10a.

hyperfine splitting, AN (1.39 mT), is consistent with trapping


of a carbon-centered radical by MNP.11,12 The additional
hyperfine splitting, AH, of 0.59 mT indicates that a proton
(nuclear spin I = 1/2) is strongly coupled to the unpaired
electron. For spin adducts of MNP, the α-proton coupling is
strongly dependent on geometry.11 This spectrum is consistent
with structure 10a that is shown in Scheme 3. The structure of
the trapped radical was further substantiated by high resolution
mass spectroscopy of the isolated product.13 Formation of this
product is consistent with pathway B (Scheme 2).
To confirm that the hyperfine coupling of 0.59 mT arises
from the proton at the α-carbon of 10a, the spin trapping
experiment was repeated with selectively deuterated substrate
9b. Spectra of the resulting spin-trapped adduct 10b are shown
in Figure 1C and 2C. Deuterium has I = 1, which splits an EPR
signal into 3 equally spaced lines, with a hyperfine coupling
constant that is a factor of 6.5 smaller than for a proton in the
same chemical environment. The much smaller splitting from
2
Figure 1. X-band EPR spectra of spin-trapped adducts in THF after
H is not resolved in the broad lines of the EPR spectrum for
workup to remove SmI2. Spectra were recorded in air with 0.54 μT B1 the air saturated solution (Figure 1C) but is partially resolved
and 110 s acquisition time. (A) Adduct 12 from reaction of 11. (B) for spectra in deoxygenated solution (Figure 2C). The
Adduct 10a from reaction of 9a. (C) Adduct 10b from reaction of 9b. hyperfine coupling constants that were obtained by simulation
To avoid back-exchange, D2O was used in the workup. of the spectra are summarized in Table 1.
10689 DOI: 10.1021/acs.joc.8b01517
J. Org. Chem. 2018, 83, 10688−10692
The Journal of Organic Chemistry Note

Table 1. Summary of Hyperfine Coupling Constants


entry compound AN (mT) AH (mT) AD (mT)
1 12 1.43 0.56
2 10a 1.39 0.59
3 10b 1.39 0.093
4 13 1.52 radical and AH = 0.56 mT, similar to the spectrum for 10a,
5 SmI2 + MNP 1.52 providing additional evidence for pathway B. When the SmI2
reaction was attempted with 2-bromoacetophenone that has
Reaction of SmI2 with MNP in the absence of either 9a or the Br on a primary carbon instead of a secondary carbon, only
9b produces the 3-line spectrum shown in Figure 3B and 3C DTBN was observed, which indicated that the additional
substituent was needed for stabilization of the resulting carbon-
centered radical. For similar substrates without Br present, only
DTBN was observed.
To rule out the possibility that reduction of the OMs group
could generate a species that impacts the radical reactions, a
spin trapping reaction was performed with SmI2 and Ms-
protected n-octanol (eq 2). Only DTBN (13) was observed.

This result is consistent with the assumption that the OMs


group is not reduced under the reaction conditions17 and acts
only as a leaving group in our previously disclosed intra-
molecular alkylation.5
The following evidence supports the assignment of pathway
B for the reactions shown in Scheme 2. (i) The carbon-
centered radicals 10a, 10b, and 12 were observed. (ii) There is
no obvious pathway by which a ketyl radical formed by
pathway A would rearrange to form the carbon centered
radical formed in pathway B. (iii) The bromine substituent is
required for the chemistry shown in Scheme 2 and for
formation of trapped radicals 10a, 10b, and 12. The only
reports of EPR spectra for spin-trapped ketyl radicals, such as
those that would be formed by pathway A, are for the small
acetone ketyl radical.18 Although the possible formation of
ketyl radicals via pathway A cannot be ruled out, the spin
trapping results support pathway B.19
Figure 3. X-band EPR spectra in THF for (A) 13, after degassing, In summary, radical trapping studies were performed for the
recorded with 0.54 μT B1 and 110 s scan time. (B) SmI2 + MNP, after SmI2 reduction of an α-bromoketone utilized previously for
degassing, recorded with 5.4 μT B1 and 13 min acquisition time. An intramolecular enolate alkylation cyclization reactions.5 EPR
off-resonance background spectrum was subtracted, which doubles
data acquisition time to 26 min. (C) SmI2 + MNP, before degassing,
analysis of the resulting trapped species support a mechanism
recorded with 8.7 μT B1 and 110 s acquisition time. in which direct reduction of the carbon−bromine bond results
in an α-centered radical rather than a ketyl radical
intermediate. While a recent study has shown some
for which AN = 1.52 mT. This hyperfine splitting is significantly experimental support for the presence of ketyl radical
larger than that for the spin trapped adducts 10a or 10b. In intermediates in similar reductions, the substrates for that
spin trapping experiments with MNP that did not produce a study contained electronically dissimilar α-bromoamides.20
radical that could be trapped, the spectrum of the relatively Further support that SmI2 reduction of α-bromoketones
stable di-tert-butylnitroxide (DTBN) 13 has been observed.7d proceeds through an α-centered radical was established
The spectrum of commercially available DTBN in THF is through spin-trapping studies using deuterium-labeled and
shown in Figure 3A and has AN = 1.52 mT. This value of AN in differentially substituted substrates. These studies should
THF is similar to reported values of AN = 1.55 mT in facilitate further development of novel carbon−carbon bond
benzene7d and 1.7114 to 1.72 mT15 in H2O. DTBN has also forming reactions through the generation of samarium
been reported as an impurity in MNP15 that can arise from enolates.


thermal or photochemical decomposition.14,16
To test the generality of the spin trapping results, the EXPERIMENTAL SECTION
reaction of SmI2 with 2-bromopropiophenone 11 was studied General Experimental Details. All starting materials were
(eq 1). The spectrum of the resulting spin-trapped adduct 12 is purchased from commercial sources. Tetrahydrofuran and dichloro-
shown in Figures 1A and 2A. The spectrum has the methane solvents were purified prior to use by an Innovative
characteristic AN = 1.43 mT of the trapped carbon-centered Technology Pur Solv system. Reactions were conducted under

10690 DOI: 10.1021/acs.joc.8b01517


J. Org. Chem. 2018, 83, 10688−10692
The Journal of Organic Chemistry Note

nitrogen atmosphere. NMR spectra were acquired on a Bruker 500 6-Hydroxy-1-phenylhexan-1-one-d2. Yield: 285.3 mg (76%). 1H
MHz NMR. Proton spectra were acquired at 500 MHz. Carbon NMR (500 MHz, Chloroform-d) δ 7.95 (d, J = 7.8 Hz, 2H), 7.55 (s,
spectra were acquired at 126 MHz. Infrared spectra were measured on 1H), 7.46 (t, J = 7.7 Hz, 2H), 3.65 (t, J = 6.5 Hz, 2H), 1.76 (m, 2H),
a Thermo Scientific IR. A benzophenone sample gave a carbonyl 1.62 (dt, J = 14.4, 6.7 Hz, 2H), 1.45 (p, J = 7.5, 6.9 Hz, 2H). HRMS
stretch at 1655 cm−1. Melting points where acquired on a DigiMelt (ESI+) m/z: [M + Li]+ Calcd for C12H14D2O2Li 201.1436; Found
MPA160 melting point instrument. Products were purified by column 201.1445.
chromatography on silica gel using a Teledyne Isco CombiFlash Rf 2-Bromo-6-hydroxy-1-phenylhexan-1-one-d. Yield: 145.7 mg
200 or by manual column chromatography. High resolution mass (37%). 1H NMR (500 MHz, Chloroform-d) δ 8.02 (d, J = 7.3 Hz,
spectra were obtained on a Waters Synapt G2 HDMS Quadrupole/ 2H), 7.61 (t, J = 7.6 Hz, 1H), 7.50 (t, J = 7.8 Hz, 2H), 3.70 (q, J = 6.1
ToF mass spectrometer with electrospray ionization (Central Hz, 3H), 3.46 (t, J = 6.7 Hz, 1H), 2.08 (m, 6H), 1.60 (m, 6H).
Analytical Laboratory, University of Colorado Boulder). Where two HRMS (ESI+) m/z: [M + Li]+ Calcd for C12H14DBrO2Li 278.0478;
molecular ion weights are reported, the second is the M+2 molecular Found 278.0482.
isotope of Br. The spin adducts 10a, 10b, and 12 are sufficiently stable 9b. Yield: 122.1 mg (65%). 1H NMR (500 MHz, Chloroform-d) δ
that an aqueous workup procedure was used to remove the SmI2 and 8.02 (d, J = 7.3 Hz, 2H), 7.62 (t, J = 7.4 Hz, 1H), 7.51 (t, J = 7.7 Hz,
side products prior to recording the EPR spectra, as described below. 2H), 4.27 (m, 3H), 3.46 (t, J = 6.3 Hz, 1H), 3.02 (s, 4H), 2.21 (m,
Some samples were also deoxygenated by freeze−pump−thaw on a 2H), 1.98 (m, 2H), 1.84 (m, 2H), 1.70 (m, 1H). HRMS (ESI+) m/z:
vacuum line, which produced narrower lines and improved resolution [M + Na]+ Calcd for C13H16DBrO4SNa 371.9991; Found 371.9986.
of the nuclear hyperfine couplings. EPR spectra were recorded on a Procedure for Spin Trapping. To a flamed dried flask that was
modified Bruker E500T using rapid scan EPR spectroscopy.21 evacuated and backfilled with N2 three times was added a [0.05 M]
Procedure for Preparation of α-Bromoketone 9a. The solution of the compound of interest (0.025 mmol, 0.5 mL) in THF
synthesis and characterization of this compound has been described and a [0.025 M] solution of MNP (0.0125 mmol, 0.5 mL) in THF,
previously.5 followed by a [0.1 M] solution of SmI2 (0.05 mmol, 0.5 mL). After
Procedure for Preparation of 6-Oxo-6-phenylhexyl Meth- mixing for a few seconds a small portion was removed and added to
anesulfonate (Desbromo 9a). To a flask charged with 6-hydroxy- an EPR tube. A sample of 10a was analyzed by HRMS.
1-phenylhexan-1-one (1.025 mmol), synthesized using previously 10a. HRMS (ESI+) m/z: [M+Li]+ Calcd for C17H26NO5SLi
reported procedure,5 dichloromethane (5.13 mL, [0.2 M]) was added. 363.1692; Found: 363.1713.
The flask was cooled to 0 °C (ice-water bath). After the flask was Procedure for Aqueous Workup of Spin Trapping Experi-
ments under Normal Atmosphere. To stirring mixture of spin
cooled, triethylamine (1.538 mmol, 1.5 equiv) and methanesulfonyl trap, SmI2, and compound of interest was added 5 mL of saturated
chloride (1.128 mmol, 1.1 equiv) were added. After an hour of stirring NaHCO3(aq) solution. This was stirred for 1 h under normal
at 0 °C, the reaction was diluted with dichloromethane (50 mL) and atmosphere. After an hour, the solution was diluted with ethyl acetate
washed with cold H2O (50 mL). The organic layer was collected, and (25 mL) and the aqueous layer was removed. The organic layer was
the aqueous layer was extracted with dichloromethane (50 mL). The washed with saturated brine solution. The organic layer was collected,
organic layers were combined, dried over anhydrous sodium sulfate, dried over anhydrous sodium sulfate, filtered, and concentrated to
filtered, and concentrated to dryness. The crude product was purified dryness. The resulting residue was dissolved in THF (3 mL). A small
by column chromatography. Yield: 212.0 mg (76%). Rf: 0.67 (1:1 portion was withdrawn and added to an EPR tube.
ethyl acetate:hexanes). mp (°C): 57.7−58.9. 1H NMR (500 MHz, Procedure for Aqueous Workup of Spin Trapping Experi-
Chloroform-d) δ 7.96 (d, J = 7.1 Hz, 2H), 7.57 (t, J = 7.4 Hz, 1H), ments under Oxygen Free Atmosphere. To stirring mixture of
7.47 (t, J = 7.7 Hz, 2H), 4.25 (t, J = 6.5 Hz, 2H), 3.02 (m, 5H), 1.81 spin trap, SmI2, and compound of interest was added 5 mL of
(m, 4H), 1.52 (m, 2H). 13C NMR (126 MHz, CDCl3) δ 199.8, 136.9, saturated NaHCO3(aq) solution. The saturated NaHCO3(aq) had N2
133.1, 128.6, 128.0, 69.8, 38.2, 37.4, 29.1, 25.2, 23.5. IR (cm−1): 3026, bubbled through the solution for at minimum 15 min prior to
2923, 2853, 1771, 1682, 1597, 1580, 1449, 1412, 1351, 1259, 1216, addition. The mixture was stirred for 1 h under N2. After an hour, the
1201, 1171, 1067, 1036, 972, 945, 814, 755, 600. HRMS (ESI+) m/z: solution was diluted with THF (3 mL) and the aqueous layer was
[M + H]+ Calcd for C13H19O4S 271.1004; Found 271.1003. removed with a syringe. The organic layer was washed with saturated
Procedure for Preparation of Octyl Methanesulfonate. 1- brine solution. The brine solution had N2 bubbled through the
Octanol (5.00 mmol) was mesylated using previously reported solution for a minimum of 15 min prior to addition. The aqueous
procedure.22 layer was removed. A small portion of the organic layer was removed
Procedure for Preparation of SmI2. Samarium metal (Sm0) and added to an EPR tube that was evacuated and backfilled with N2
(2.60 mmol, 1.3 equiv), 1,2-diiodoethane (2.00 mmol, 1 equiv), and three times.


THF (20 mL, [0.1 M]) were added to a flask that had been flame-
dried, evacuated, and backfilled with N2 atmosphere. The reaction was ASSOCIATED CONTENT
sonicated under positive N2 atmosphere for 10 min. After which, the
solution of SmI2 was continuous stirred until ready for use. *
S Supporting Information
Procedure for Preparation of NaOD. Sodium hydroxide The Supporting Information is available free of charge on the
(NaOH) (1.00 mmol) was dissolved in 10 mL of D2O. The solution ACS Publications website at DOI: 10.1021/acs.joc.8b01517.
was boiled until less than 1 mL of solvent remained. After which, the
solution was cooled to 25 °C, and 10 mL of D2O was added and NMR spectra for all new compounds (PDF)


subjected to boiling until almost dry again. This process was repeated
two additional times. After the final boiling, the volume was brought
to 10 mL with D2O. AUTHOR INFORMATION
Procedure for α-Hydrogen−Deuterium Exchange. To a flask Corresponding Author
charged with 6-hydroxy-1-phenylhexan-1-one (369.8 mg, 1.924
mmol) substrate was added 1 mL of the [0.1 M] NaOD in D2O
*E-mail: bryan.cowen@du.edu.
solution followed by 9 mL of D2O. The reaction was refluxed at 100 ORCID
°C for 8 h. This was sufficient to undergo hydrogen−deuterium Christopher D. Aretz: 0000-0002-7125-4388
exchange on the α-carbon of the substrate. The crude product was α- Joseph E. McPeak: 0000-0001-8677-6405
brominated and then mesylated using previously reported procedure.5
This exchange was confirmed by 1H NMR and HRMS. In the 1H Gareth R. Eaton: 0000-0001-7429-8469
NMR spectrum the resonances associated with the α-proton were Sandra S. Eaton: 0000-0002-2731-7986
absent. Bryan J. Cowen: 0000-0001-8543-3195
10691 DOI: 10.1021/acs.joc.8b01517
J. Org. Chem. 2018, 83, 10688−10692
The Journal of Organic Chemistry Note

Author Contributions (14) Spulber, M.; Schlick, S. Fragmentation of Perfluorinated


C.D.A and J.E.M. contributed equally. Membranes Used in Fuel Cells: Detecting Very Early Events by
Selective Encapsulation of Short-Lived Fragments in β-Cyclodextrin.
Notes J. Phys. Chem. B 2011, 115, 12415−12421.
The authors declare no competing financial interest. (15) Alvarez, B.; Demicheli, V.; Duran, R.; Trujillo, M.;


Cevrenansky, C.; Freeman, B. A.; Radi, R. Inactivation of Human
Cu,Zn Superoxide Dismutase by Peroxynitrite and Formation of
ACKNOWLEDGMENTS Histidinyl Radical. Free Radical Biol. Med. 2004, 37, 813−822.
Partial support of this work by NSF CHE-1227992 (G.R.E. (16) Perkins, M. J. In Adv. Phys. Org. Chem., Vol. 17; Gold, V.,
and S.S.E.) and NIH CA177744 (G.R.E. and S.S.E.) and by the Bethell, D., Eds.; Academic Press: 1980; p 1.
University of Denver is gratefully acknowledged. We also thank (17) Alkyl tosylates were reduced using Sml2 but only in the
Dr. Thomas Lee and Dr. Danijel Djukovic of the University of presence of Lewis basic additives: (a) Ankner, T.; Hilmersson, G.
Instantaneous Deprotection of Tosylamides and Esters with SmI2/
Colorado Boulder for the acquisition of HRMS data.


Amine/Water. Org. Lett. 2009, 11, 503−506. (b) Szostak, M.; Spain,
M.; Procter, D. J. Recent advances in the chemoselective reduction of
REFERENCES functional groups mediated by samarium(II) iodide: a single electron
(1) (a) Kagan, H. B. Twenty-five years of organic chemistry with transfer approach. Chem. Soc. Rev. 2013, 42, 9155−9183.
diiodosamarium: an overview. Tetrahedron 2003, 59, 10351−10372. (18) Chiu, T.-M.; Siemiarczuk, A.; Wong, S. K.; Bolton, J. R. Time
Resolution Enhancement Technique Applied to a Study of the
(b) Procter, D. J.; Flowers, R. A.; Skrydstrup, T. Organic Synthesis
Absolute Rate of Reaction of Ketyl Radicals with a Spin Trap Using
Using Samarium Diiodide: A Practical Guide; Royal Society of
Flash Photolysis Electron Paramagnetic Resonance. J. Phys. Chem.
Chemistry Publishing: Cambridge, 2010. (c) Szostak, M.;
1985, 89, 3343−3347.
Fazakerley, N. J.; Parmar, D.; Procter, D. J. Cross-Coupling Reactions
(19) We are unable to quantify the formation of spin trapped
Using Samarium(II) Iodide. Chem. Rev. 2014, 114, 5959−6039.
adducts 10a, 10b, or 12.
(2) Rudkin, I. M.; Miller, L. C.; Procter, D. J. Samarium enolates and
(20) Yang, S.; Xi, Y.; Chen, J.; Yang, Z. Mechanistic Study of SmI2-
their application in organic synthesis. Organomet. Chem. 2008, 34,
Mediated Reformatsky Reaction for Macrolactam Formation Using a
19−45.
Cyclopropyl Group as a Probe. Isr. J. Chem. 2017, 57, 331−334.
(3) Szostak, M.; Procter, D. J. Beyond Samarium Diiodide: Vistas in
(21) Eaton, S. S.; Quine, R. W.; Tseitlin, M.; Mitchell, D. G.; Rinard,
Reductive Chemistry Mediated by Lanthanides(II). Angew. Chem., Int.
G. A.; Eaton, G. R. In Multifrequency Electron Paramagnetic Resonance:
Ed. 2012, 51, 9238−9256.
Data and Techniques; Misra, S. K., Ed.; Wiley: 2014; p 3.
(4) Prasad, E.; Flowers, R. A. Mechanistic Study of β-Substituent
(22) Hong, Y.-R.; Gorman, C. B. Synthetic Approaches to an
Effects on the Mechanism of Ketone Reduction by SmI2. J. Am. Chem.
Isostructural Series of Redox-Active, Metal Tris(bipyridine) Core
Soc. 2002, 124, 6357−6361.
Dendrimers. J. Org. Chem. 2003, 68, 9019−9025.
(5) Aretz, C. D.; Escobedo, H.; Cowen, B. J. Cyclopentane
Formation from Flexible Precursors Using Samarium(II) Reagents.
Eur. J. Org. Chem. 2018, 2018, 1880−1884.
(6) Molander, G. A.; Hahn, G. Lanthanides in Organic Synthesis. 2.
Reduction of α-Heterosubstituted Ketones. J. Org. Chem. 1986, 51,
1135−1138.
(7) For examples of the use of spin traps to study reactions involving
SmI2, see: (a) Ma, S.; Lu, X. Studies on Samarium Diiodide Initiated
Addition Reaction of Fluoroalkyl Iodides to Alkynes and its
Mechamism. Tetrahedron 1990, 46, 357−364. (b) Tamura, R.;
Susuki, S.; Azuma, N.; Matsumoto, A.; Toda, F.; Ishii, Y. A New
Synthesis of α-Asymmetric Nitroxide Radicals. J. Org. Chem. 1995, 60,
6820−6825. (c) Katritzky, A.; He, H.-Y.; Qiu, G.; Bratt, P. J.; Parrish,
S. H., Jr.; Angerhofer, A. EPR Studies on the SmI2-Promoted
Coupling of N-(N′,N′-Dialkylaminoalkyl)benzotriazoles. Org. Lett.
1999, 1, 1755−1757. (d) Alberti, A.; Benaglia, M.; Macciantelli, D.
Mechanistic Studies of Radical-Based Processes. Use and Misuse of
EPR Spectroscopy. Org. Lett. 2000, 2, 1553−1555. (e) Sono, M.;
Hanamura, S.; Furumaki, M.; Murai, H.; Tori, M. First Direct
Evidence of Radical Intermediates in Samarium Diiodide Induced
Cyclization by ESR Spectra. Org. Lett. 2011, 13, 5720−5723.
(8) Janzen, E. G.; Haire, D. L. In Two Decades of Spin Trapping.
Advances in Free Radical Chemistry, Vol. 1; JAI Press Inc.: 1990; p 253.
(9) Kemp, T. Kinetic Aspects of Spin Trapping. Prog. React. Kinet.
Mech. 1999, 24, 287−358.
(10) (a) Hasegawa, E.; Curran, D. P. Rate Constants for the
Reactions of Primary Alkyl Radicals with SmI2 in THF/HMPA.
Tetrahedron Lett. 1993, 34, 1717−1720. (b) Shabangi, M.; Kuhlman,
M. L.; Flowers, R. A. Mechanism of Reduction of Primary Alkyl
Radicals by SmI2−HMPA. Org. Lett. 1999, 1, 2133−2135.
(11) Madden, P.; Taniguchi, H. An in Situ Radiolysis Time-
Resolved Electron Spin Resonance Study of 2-Methyl-2-nitro-
sopropane Spin Trapping Kinetics. J. Am. Chem. Soc. 1991, 113,
5541−5547.
(12) Buettner, G. R. Spin Trapping: ESR Parameters of Spin
Adducts. Free Radical Biol. Med. 1987, 3, 259−303.
(13) See Experimental Section for details.

10692 DOI: 10.1021/acs.joc.8b01517


J. Org. Chem. 2018, 83, 10688−10692

You might also like