You are on page 1of 6

Note

Cite This: J. Org. Chem. XXXX, XXX, XXX−XXX pubs.acs.org/joc

Synthesis of Nitriles from Primary Amides or Aldoximes under


Conditions of a Catalytic Swern Oxidation
Rui Ding, Yongguo Liu, Mengru Han, Wenyi Jiao, Jiaqi Li, Hongyu Tian,* and Baoguo Sun*
Beijing Advanced Innovation Center for Food Nutrition and Human Health, Beijing Key Laboratory of Flavor Chemistry, Beijing
Technology and Business University, Beijing 100048, China
*
S Supporting Information

ABSTRACT: The preparation of nitriles from primary amides or


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

aldoximes was achieved by using oxalyl chloride with a catalytic amount


of dimethyl sulfoxide in the presence of Et3N. The reactions were complete
Downloaded via KAOHSIUNG MEDICAL UNIV on October 7, 2018 at 10:55:57 (UTC).

within 1 h after addition at room temperature. A diverse range of cyano


compounds were obtained in good to excellent yields, including aromatic,
heteroaromatic, cyclic, and acyclic aliphatic species.

N itriles are important organic compounds either as


versatile precursors in synthesis or as valuable end
products. The cyano group can be easily converted to other
sulfenllactonization10 and chlorolactonization of alkenoic
acids,11 bromination of alkenes, alkynes, and ketones,12 and
preparation of α,β-unsaturated γ-lactones.13 We noticed that
functional groups, such as carboxyl derivatives, amines, primary amides could be dehydrated to afford nitriles under
ketones, and various heterocycles.1 Many cyanated compounds Swern oxidation conditions; however, DMSO and (COCl)2
display attractive applications as pharmaceuticals, agrochem- were used in stoichiometric amounts and the reactions needed
icals, or fragrances.2 A plethora of methods for the synthesis of to be carried out at −78 °C.9 Obviously, it is very inconvenient
nitriles has been developed over the past century, which can be to run reactions at such a low temperature. The work of
divided into two classes, one involving introduction of a cyano Malkov and Rubtsov6b inspired us to investigate if the
group with inorganic or organic cyanating reagents3 and the dehydration of amides or aldoximes may be achieved under
other involving creation of a cyano group by chemical conditions of a catalytic Swern oxidation at room temperature.
transformation of other functional groups.4 Herein, we report the successful results of the dehydration of
The dehydration of amides or aldoximes is the most amides or aldoximes to afford nitriles with (COCl)2 and a
commonly used approach for the creation of a cyano group by catalytic amount of DMSO in the presence of Et3N at room
far. A wide variety of reagents have been described for the temperature (Scheme 1).
transformation of primary amides or aldoximes to nitriles.5
However, many of these existing methods suffer from Scheme 1. Dehydration of Amides or Aldoximes to Nitriles
drawbacks such as drastic reaction conditions, limited substrate by a Catalytic Swern Oxidation
scope, not readily available reagents, poor yields, and long
reaction time. Several attractive methods have been reported
recently in which the dehydration of primary amides or
aldoximes could be achieved via catalysis under mild
conditions.6 It is worth highlighting that a highly efficient
catalytic Appel-type dehydration of amides to nitriles has been
developed by Malkov and Rubtsov using oxalyl chloride and
triethylamine along with triphenylphosphine oxide as a
catalyst.6b The investigation began by employing benzamide (1a) as a
The combination of DMSO and oxalyl chloride has been model substrate (Table 1). First, the experiment was carried
used widely in the oxidation of primary and secondary alcohols out under conditions mimicking those in the protocol from
since Swern first reported this method in 1978.7 In addition, it amides to nitriles developed by Malkov and Rubtsov,6b in
was also very efficient for other transformations, such as which 2 equiv of oxalyl chloride in anhydrous MeCN was
methylthiomethyl esterification of carboxylic acids8 and added dropwise to the solution of 1a, 1 mol % of DMSO, and
dehydration of primary amides.9 In our recent work, novel
application of the combined reagents of R2SO (R = Me or Ph) Received: August 24, 2018
and (COX)2 (X = Cl or Br) has been developed, such as Published: September 21, 2018

© XXXX American Chemical Society A DOI: 10.1021/acs.joc.8b02190


J. Org. Chem. XXXX, XXX, XXX−XXX
The Journal of Organic Chemistry Note

Table 1. Optimization of the Reaction Conditions of Table 2. Dehydration of Amides to Nitriles under the
Dehydration of Amides Conditions of a Catalytic Swern Oxidation

DMSO (COCl)2 yield


entry (equiv) (equiv) base (equiv) time (%)
1 0.01 2.0 Et3N (3.0) 10 min 92a
2 0.01 2.0 0 24 h 1b
3 0.01 2.0 Et3N (3.0) 24 h 1b,c
4 0.01 1.2 Et3N (2.5) 40 min 90a
5 0 1.2 Et3N (2.5) 24 h 0b
6 0.01 1.2 DBU (2.5) 24 h 0b
7 0.01 1.2 NMM (2.5) 24 h 58b
a
Isolation yield after chromatographic purification. bDetermined by
GCMS. cThe only difference against the conditions of entry 1 was
that Et3N was added last.

3 equiv of Et3N in anhydrous MeCN at rt (entry 1). The


reaction was complete in 10 min to produce benzonitrile (2a)
almost quantitatively. A reaction was also carried out in the
absence of Et3N, in which only a trace of benzonitrile (2a) was
detected by GCMS even after 24 h of stirring at rt (entry 2).
Another experiment was performed under conditions similar to
those in entry 1 except that 3 equiv of Et3N was added to the
reaction mixture at last (entry 3). The conversion of
benzonitrile (2a) was only 1% after 24 h of stirring at rt, In addition, aliphatic linear (1i, 1j, 1k), cyclic (1l), and
which was similar to that in entry 2. These results indicated branched (1m, 1n) amides were dehydrated to produce the
that it was feasible that the dehydration of an amide was corresponding nitriles 2i−n in 80−87% yields, which were
carried out under the conditions of a catalytic Swern oxidation slightly lower than those from the above unsaturated amides.
at rt, in which Et3N was needed to be added to the reaction Phenyl-protected α-hydroxy amide 1k was also a compatible
mixture in advance. One more experiment was carried out to substrate, which gave the desired nitrile 2k in 80% yield. The
optimize the amounts of (COCl)2 and Et3N, in which they same protocol also was applied to the two α-amino amides 1o
were reduced to 1.2 and 2.5 equiv, respectively (entry 4). It and 1p with Boc and Cbz protective groups, respectively, and
was observed that the reaction proceeded slightly slower than the corresponding nitriles 2o and 2p were obtained in 83% and
that in entry 1 but produced the desired nitrile 2a in a very 76% yields, respectively, with the protective groups intact.
close yield. The reaction was also attempted in the absence of Moreover, the specific rotation values of the products indicated
DMSO, but no product was detected (entry 5). This result that no racemization occurred during the dehydration process.
indicated that DMSO is indispensable for the reaction, which Most of the substrates included in Table 2 were also
is consistent with that reported by Nakajima and Ubukata.9 In investigated in the work of Malkov and Rubtsov.6b For these
addition, two more bases were examined to see if they could substrates, the yields obtained in our work are comparable with
promote this reaction. No product was obtained in the those obtained via a catalytic Appel reaction. In addition, an
presence of DBU (entry 6), whereas the reaction gave the acid-sensitive Boc-protected substrate 1o was converted to the
desired nitrile in 58% yield in the presence of NMM (entry 7). corresponding nitrile in a good yield in our work, whereas it
In comparison, Et3N is the most effective base for this was mentioned that Boc-protected amino acids could not be
conversion. converted to the desired nitriles under the conditions of a
The dehydration of more amides was explored under the catalytic Appel reaction.6b
same conditions as those in entry 4 of Table 1. The results are Among the substrates mentioned above, some were also
shown in Table 2. Three benzamides with substituted groups examined by Nakajima and Ubukata9 under Swern oxidation
on the phenyl ring (1b−d) were converted to the conditions with stoichiometric use of DMSO, including 1c, 1d,
corresponding benzonitriles (2b−d) in good to excellent 1g, 1h, 1l, and 1p. Except for p-methoxybenzamide 1c with a
yields, regardless of whether they bore electron-withdrawing similar yield and 3,5-dinitrobenzamide 1d with a lower yield,
group (2b, 2d) or electron-donating group (2c). Among the the other amides were converted to the corresponding nitriles
aromatic amides, 1d with two electron-withdrawing nitro in higher yields in our work. In the work of Nakajima and
groups on the phenyl ring afforded the nitrile in the lowest Ubukata,9 all of the reactions were carried out at −78 or −78
yield of 80%. Two heterocyclic amides 1e and 1f were then °C to rt with the combination of stoichiometric amounts of
investigated. 2-Cyanothiophene 2e and 3-cyanopyridine 2f (COCl)2, DMSO, and Et3N, in which the addition of Et3N
were obtained in 89% and 90% yields, respectively. The could be done either with (COCl)2 at the same time or at last
dehydration of 2-naphthamide 1h gave 2-cyanonaphthalene 2h without any effect on yields. In contrast, the reactions were
in a high yield of 96%. Likewise, the unsaturated cinnamamide able to be performed at rt mildly due to the usage of catalytic
1g was also converted smoothly to the nitrile 2g in 92% yield. amount of DMSO in our present work. It was noteworthy that
B DOI: 10.1021/acs.joc.8b02190
J. Org. Chem. XXXX, XXX, XXX−XXX
The Journal of Organic Chemistry Note

it was necessary to add Et3N to the reaction mixture in advance Table 3. Optimization of the Reaction Conditions of
of our protocol, which indicatd that Et3N acted as base to Dehydration of Aldoximes
eliminate the intermediates to release DMSO required for the
next reaction cycle. The possible mechanistic pathway for the
dehydration of amides under catalytic Swern oxidation
conditions was proposed as follows (Scheme 2). DMSO (A)

Scheme 2. Possible Mechanistic Pathway for the


DMSO (COCl)2 yield
Dehydration of Amides under Catalytic Swern Oxidation entry (equiv) (equiv) base (equiv) time (%)
Conditions 1 0 1.5 0 24 h 0a
2 0.01 1.2 0 30 min 2a
3 0.05 1.5 0 24 h 100a
4 0.05 1.5 TEA (3.0) 20 min 100a
5 0.01 1.2 TEA (2.5) 1h 93b
6 0.01 1.2 DBU (2.5) 24 h 0a
7 0.01 1.2 NMM (2.5) 2h 100a
a b
Determined by HNMR. Isolation yield after chromatographic
purification.

and 1.5 equiv of (COCl)2 in the absence of base (entry 3). The
addition of Et3N (3.0 equiv) reduced the reaction time from
24 h to 20 min (entry 4). The amounts of reagents were
optimized further. The reaction was complete in 1 h an
reacted with (COCl)2 to produce the intermediate chlor- produced 2a in 93% isolated yield in the presence of 0.01 equiv
odimethylsulfonium chloride (B), which was attacked by the of DMSO, 1.2 equiv of (COCl)2, and 2.5 equiv of Et3N (entry
oxygen of an amide through a nucleophilic substitution in the 5). DBU and NMM were also used in attempts to replace Et3N
presence of Et3N to give the intermediate C. It was then to promote the dehydration of 3a. No nitrile was obtained in
deprotonated by Et3N to generate the intermediate D, which the presence of DBU (entry 6). In contrast, the nitrile was
underwent elimination to afford the corresponding nitrile and obtained in the presence of NMM in a yield comparable to
release DMSO (A) to enter the next cycle. No product was that with Et3N but over a longer reaction time of 2 h (entry 7).
obtained when Et3N was replaced by DBU, which might be These results were similar to those obtained from the
due to the reaction between DBU and (COCl)2.14 In contrast, dehydration of amides.
NMM gave the desired product in a lower yield for a longer Five other aldoximes underwent dehydration under the
reaction time than Et3N, which indicated that the elimination above optimized conditions: o-bromobenzaldehyde oxime 3b,
of the intermediate C to the nitrile is the rate-limiting step thiophene-2-carbaldehyde oxime 3e, trans-cinnamaldehyde
influenced by the basicity of the amine. oxime 3g, cyclohexanecarboxaldehyde oxime 3l, and 3-
There is an alternative possible pathway for the trans- phenylpropanal oxime 3q. The results are shown in Table 4.
formation of the intermediate C to nitrile; i.e., the All of these substrates were converted to the corresponding
deprotonation may occur on the methyl group to give the nitriles in high yields, and the reactions were complete in 1 h.
intermediate E, which undergoes a retro-heteroene reaction to Yadav et al. reported that bromodimethylsulfonium bromide
generate nitrile. In order to clarify the mechanism, DMSO-d6 was used to convert aldoximes and primary amides to nitriles,5a
was used for the dehydration of benzamide (1a), and the in which it was claimed that the presence of a base slowed the
reaction mixture was analyzed by GC−EIMS after the reaction
was complete. Compared with the mass spectrum of DMSO- Table 4. Dehydration of Aldoximes to Nitriles under the
d6, the major fragment ion peaks in the spectrum of DMSO in Conditions of a Catalytic Swern Oxidation
the reaction mixture are similar to those of the standard
DMSO-d6. However, two additional minor peaks at m/z 83
and 63 both with an intensity of 2.6% appear in the mass
spectrum of DMSO in the reaction mixture, which indicates
that a small amount of DMSO-d6 was converted into DMSO-
d5 after the reaction. These results indicate that the
transformation of the intermediate C to nitrile occurs mainly
by an E2 mechanism but do not exclude a retro-heteroene
mechanism to some extent.
The application scope of the present method was further
explored by transforming aldoximes to the nitriles. The
dehydration of benzaldehyde oxime (3a) was investigated
under different conditions in order to optimize reaction
conditions (Table 3). The reaction failed to produce
benzonitrile (2a) in the presence of (COCl)2 only if without
DMSO (entry 1). Compound 3a was converted into 2a
thoroughly after 24 h by treatment with 0.05 equiv of DMSO
C DOI: 10.1021/acs.joc.8b02190
J. Org. Chem. XXXX, XXX, XXX−XXX
The Journal of Organic Chemistry Note

dehydration of aldoximes. In contrast, it was observed that the with a catalytic amount of DMSO. This approach has several
reactions of all the aldoximes in our work were very slow in the advantages, such as simple operation, mild conditions, high
absence of Et3N and completed in 5−48 h. The possible yields, and short reaction time. It offers a practical alternative
mechanistic pathway for the dehydration of aldoximes is shown for the preparation of nitriles by the dehydration of primary
in Scheme 3, which is similar to the catalytic cycle above for amides or aldoximes. The investigation about the generality of
this method is still ongoing in our laboratory.
Scheme 3. Possible Mechanistic Pathway for the
Dehydration of Aldoximes under Catalytic Swern Oxidation
Conditions
■ EXPERIMENTAL SECTION
General Information. NMR spectra were obtained on a Bruker
AV 300 spectrometer (1H NMR at 300 MHz, 13C NMR at 75 MHz)
in CDCl3 using TMS as an internal standard. Chemical shifts (δ) are
given in ppm and coupling constants (J) in hertz. GC−MS data were
obtained on an Agilent 7890B-5977A under the following conditions:
capillary column HP-5MS 5% Phenyl Methyl Silox (30 m × 0.25 mm
× 0.25 um); oven temperature programmed from 50 to 150 °C at a
rate of 15 °C/min, then 10 to 280 °C at a rate of 10 °C/min; carrier
gas, helium; flow rate, 1.2 mL/min; electron ionization, 70 eV; ion
source temperature, 230 °C. TLC was performed with precoated TLC
plates, silica gel 60F-254, layer thickness 0.25 mm. Flash
chromatography separations were performed on 200−300 mesh silica
gel. Reagents and solvents are commercial grade and were used as
supplied. Amides (1a−p) and aldoximes (3a, 3b, 3e, and 3g) are
commercially available and were purchased from Innochem; the other
aldoximes (3l and 3q) were synthesized in our laboratory.
the dehydration of amides. The presence of Et3N is crucial for General Procedure for Dehydration of Primary Amides or
Aldoximes to Nitriles. Amide 1 (3 mmol, 1.0 equiv) or aldoxime 3
the regeneration of DMSO, which is necessary for the next (3 mmol, 1.0 equiv) was dissolved in 10 mL of anhydrous acetonitrile,
reaction cycle. followed by addition of DMSO (2.5 mg, 0.03 mmol, 0.01 equiv) and
In work of Denton et al. about the dehydration of oximes Et3N (1.04 mL, 7.5 mmol, 2.5 equiv). Oxalyl chloride (0.31 mL, 3.6
using oxalyl chloride in combination with 5 mol % of Ph3PO, a mmol, 1.2 equiv) in anhydrous acetonitrile (5 mL) was added
catalytic cycle was proposed involving the activated mono dropwise at 20 °C and stirred at room temperature. After completion
oxime ester generated from the reaction of oxime and oxalyl of reaction as indicated by TLC or GCMS, the mixture was filtered
chloride as intermediate, which was then converted to nitrile and concentrated in vacuo. Distilled water (15 mL) was added, the
by Ph3PO or chlorophosphonium salt. In order to explore the mixture was extracted with EtOAc (3 × 10 mL), and combined
extracts were washed with brine (2 × 20 mL), dried over anhydrous
possibility of the similar pathway, two more experiments were
Na2SO4, filtered, and concentrated in vacuo. Purification by flash
carried out, in which benzamide 1a or benzaldehyde oxime 3a chromatography (silica gel, petroleum ether/ethyl acetate = 9:1)
reacted with 1.5 equiv of oxalyl chloride first for 5 h at room afforded the corresponding nitrile 2.
temperature, and then 0.05 equiv of DMSO was added. In the General Procedure for Synthesis of Aldoximes from
case of benzamide 1a, no product was detected after 24 h. In Aldehydes. To a mixture of an aldehyde (10 mmol, 1.0 equiv)
contrast, benzaldehyde oxime 3a was converted to the desired and hydroxylamine hydrochloride (1.38 g, 20 mmol, 2.0 equiv) in
nitrile in 90% yield after stirring for 6 h. These results indicated anhydrous dichloromethane (30 mL) was added pyridine (3.22 mL,
that for oximes, a catalytic cycle involving the activated mono 40 mmol, 4.0 equiv). The mixture was stirred for 24 h at room
temperature. HCl (aq 2.0 M, 30 mL) was added to the reaction
oxime ester (G) was possible (Scheme 4). However, in the
mixture, and the solution was extracted with dichloromethane (3 × 30
case of amides, the mechanism shown in Scheme 2 is more mL). The combined extracts were washed with brine (50 mL), dried
reasonable. over anhydrous Na2SO4, filtered, and concentrated in vacuum to give
In conclusion, a highly efficient preparation method of the corresponding aldoxime 3.
nitriles from primary amides or aldoximes has been developed Analytical Data. Benzonitrile (2a).6b Colorless oil, 285 mg, 92%
using the combination of oxalyl chloride and triethylamine yield. 1H NMR (300 MHz, CDCl3): δ 7.68−7.64 (m, 2H, H-o-
phenyl), 7.64−7.58 (m, 1H, H-p-phenyl), 7.50−7.44 (m, 2H, H-m-
Scheme 4. Possible Mechanistic Pathway for the phenyl). 13C NMR (75 MHz, CDCl3): δ 132.9 (C-p-phenyl), 132.3
Dehydration of Aldoximes through the Activated Mono (C-o-phenyl), 129.2 (C-m-phenyl), 119.0 (C−CN), 112.6 (C1-
phenyl).
Oxime Ester 2-Bromobenzonitrile (2b).15 White solid, 516 mg, 95% yield. 1H
NMR (300 MHz, CDCl3): δ 7.70−7.64 (m, 2H, H−C3-phenyl and
H−C6-phenyl), 7.49−7.39 (m, 2H, H−C4-phenyl and H−C5-
phenyl). 13C NMR (75 MHz, CDCl3): δ 134.4 (C6-phenyl), 134.0
(C4-phenyl), 133.3 (C3-phenyl), 127.8 (C5-phenyl), 125.4 (C2-
phenyl), 117.2 (C−CN), 116.0 (C1-phenyl). The NMR peak
assignments were confirmed by the HMQC and HMBC spectra.
4-Methoxybenzonitrile (2c).6b White solid, 380 mg, 95% yield. 1H
NMR (300 MHz, CDCl3): δ 7.61−7.56 (m, 2H, H-o-phenyl), 6.97−
6.93 (d, 2H, H-m-phenyl), 3.86 (s, 3H, H−OCH3). 13C NMR (75
MHz, CDCl3): δ 163.0 (C-p-phenyl), 134.1 (C-o-phenyl), 119.4 (C−
CN), 114.9 (C-m-phenyl), 104.2 (C1-phenyl), 55.7 (C-OCH3).
3,5-Dinitrobenzonitrile (2d). White solid, 463 mg, 80% yield. 1H
NMR (300 MHz, CDCl3): δ 9.27 (t, J = 2.1 Hz, 1H, H-p-phenyl),
8.85 (d, J = 2.1 Hz, 2H, H-o-phenyl). 13C NMR (75 MHz, CDCl3): δ

D DOI: 10.1021/acs.joc.8b02190
J. Org. Chem. XXXX, XXX, XXX−XXX
The Journal of Organic Chemistry Note

149.0 (C-m-phenyl), 132.1 (C-o-phenyl), 122.8 (C-p-phenyl), 116.0 CH(CH3)2). 13C NMR (75 MHz, CDCl3) δ 154.7 (CO), 118.1
(C−CN), 114.7 (C1-phenyl). (CN), 80.7 (C−OC(CH3)3), 48.4 (C-CHCN), 31.7 (C-CH(CH3)2),
Thiophene-2-carbonitrile (2e).6b Colorless oil, 292 mg, 89% yield. 28.1 (C-OC(CH3)3), 18.4 (C−CH(CH3)2), 17.9 (C−CH(CH3)2).
1
H NMR (300 MHz, CDCl3): δ 7.65−7.63 (m, 1H, H−C-3), 7.61 The NMR peak assignments were confirmed by the HMQC
(dd, J = 5.1, 1.2 Hz, 1H, H−C-5), 7.14 (dd, J = 5.1, 3.6 Hz, 1H, H− spectrum. [α]23D = −66.3 (c 1.41, MeOH).
C-4). 13C NMR (75 MHz, CDCl3): δ 137.5 (C-3), 132.7 (C-5), 127.7 (S)-Benzyl 1-Cyano-2-methylpropylcarbamate (2p).19 White
(C-4), 114.3 (C−CN), 110.0 (C-2). The NMR peak assignments solid, 530 mg, 76% yield. 1H NMR (300 MHz, CDCl3, 40 °C): δ
were confirmed by the HMQC and HMBC spectra. 7.41−7.28 (m, 5H, H-phenyl), 5.29 (br d, J = 8.4 Hz, 1H, H-NH),
Nicotinonitrile (2f).6b White solid, 281 mg, 90% yield. 1H NMR 5.15 (s, 2H, H−CH2Ph), 4.49 (br s, 1H, H−CHCN), 2.11−1.96 (m,
(300 MHz, CDCl3): δ 8.89 (d, J = 1.2 Hz, 1H, H−C-2), 8.82 (dd, J = 1H, H−CH(CH3)2), 1.08 (d, J = 6.6 Hz, 3H, H−CH(CH3)2), 1.06
4.8, 1.8 Hz, 1H, H−C-6), 7.96 (dt, J = 7.8, 1.8 Hz, 1H, H−C-4), 7.44 (d, J = 6.9 Hz, 3H, H−CH(CH3)2).13C NMR (75 MHz, CDCl3): δ
(ddd, J = 7.8, 4.8, 0.9 Hz, 1H, H−C-5). 13C NMR (75 MHz, CDCl3): 155.4 (CO), 135.7 (C1-phenyl), 128.7, 128.6, 128.3 (C2−C6-
δ 153.1 (C-6), 152.6 (C-2), 139.3 (C-4), 123.7 (C-5), 116.6 (C− phenyl), 117.8 (CN), 67.8 (C-CH2Ph), 49.1 (C-CHCN), 31.8 (C-
CN), 110.3 (C-3). The NMR peak assignments were confirmed by CH(CH3)2), 18.6 (C−CH(CH3)2), 18.0 (C−CH(CH3)2). The NMR
the HMQC spectrum. peak assignments were confirmed by the HMQC spectrum. [α]23D =
Cinnamonitrile (2g).6b Colorless oil, 357 mg, 92% yield. 1H NMR −53.8 (c 1.05, CH2Cl2).
(300 MHz, CDCl3): δ 7.48−7.37 (m, 6H, H-phenyl and H−CHPh), 3-Phenylpropanenitrile (2q).20 Colorless oil, 355 mg, 90% yield.
5.88 (d, J = 16.8 Hz, 1H, H−CHCN). 13C NMR (75 MHz, CDCl3): 1
H NMR (300 MHz, CDCl3): δ 7.37−7.20 (m, 5H, H-phenyl), 2.94
δ 150.7 (C-CHPh), 133.6 (C1-phenyl), 131.3 (C-p-phenyl), 129.2 (t, J = 7.5 Hz, 2H, H−C-3), 2.60 (t, J = 7.5 Hz, 2H, H−C-2). 13C
(C-m-phenyl), 127.4 (C-o-phenyl), 118.2 (C−CN), 96.4 (C-CHCN). NMR (75 MHz, CDCl3): δ 138.1 (C1-phenyl), 128.9 (C-m-phenyl),
2-Naphthonitrile (2h).15 White solid, 441 mg, 96% yield. 1H NMR 128.3 (C-o-phenyl), 127.3 (C-p-phenyl), 119.2 (C−CN), 31.6 (C-3),
(300 MHz, CDCl3): δ 8.24 (s, 1H, H−C-1), 7.94−7.88 (m, 3H, H− 19.4 (C-2). The NMR peak assignments were confirmed by the
C-4, H−C-5 and H−C-8), 7.69−7.57 (m, 3H, H−C-3, H−C-6 and HMQC and HMBC spectra.
H−C-7). 13C NMR (75 MHz, CDCl3): δ 134.8 (C-4a), 134.3 (C-1), Cyclohexanecarbaldehyde Oxime (3l).21 Colorless oil, 1.05 g,
132.4 (C-8a), 129.3, 129.2, 128.6, 128.2, 127.8, 126.5 (C-3, C-4, C-5, 83% yield. 1H NMR (300 MHz, CDCl3): δ 8.94 (br s, 1H, H−OH,
C-6, C-7 and C-8), 119.4 (C−CN), 109.6 (C-2). major and minor), 7.32 (d, J = 6.0 Hz, 1H, H−CHN, major), 6.53
Hexanenitrile (2i).16 Colorless oil, 233 mg, 80% yield. 1H NMR (d, J = 7.2 Hz, 1H, H−CHN, minor), 3.02−2.90 (m, 1H, H−C1-
(300 MHz, CDCl3): δ 2.33 (t, J = 7.2 Hz, 2H, H−C-2), 1.71−1.62 cyclohexane, minor), 3.26−2.15 (m, 1H, H−C1-cyclohexane, major),
(m, 2H, H−C-3), 1.49−1.32 (m, 4H, H−C-4 and H−C-5), 0.92 (t, J 1.85−1.09 (m, 10H, H-5CH2, major and minor). 13C NMR (75 MHz,
= 7.2 Hz, 3H, H−C-6). 13C NMR (75 MHz, CDCl3): δ 120.0 (C− CDCl3): δ 156.6 (C−CHN, minor), 156.1 (C−CHN, major),
CN), 30.9 (C-4), 25.2 (C-3), 22.0 (C-5), 17.2 (C-2), 13.8 (C-6). The 38.6 (C1-cyclohexane, major), 33.9 (C1-cyclohexane, minor), 30.3
NMR peak assignments were confirmed by the HMQC spectrum. (C2- and C6-cyclohexane, major), 29.5 (C2- and C6-cyclohexane,
2-Phenylacetonitrile (2j).6b Light yellow oil, 291 mg, 83% yield. minor), 26.01 (C4-cyclohexane, minor), 25.96 (C4-cyclohexane,
1
H NMR (300 MHz, CDCl3): δ 7.42−7.31 (m, 5H, H-phenyl), 3.75 major), 25.5 (C3- and C5-cyclohexane, major), 25.3 (C3- and C5-
(s, 2H, H−CH2). 13C NMR (75 MHz, CDCl3): δ 130.0 (C1-phenyl), cyclohexane, minor). The NMR peak assignments were confirmed by
129.3 (C-o-phenyl), 128.2 (C-p-phenyl), 128.0 (C-m-phenyl), 118.0 the HMQC spectrum.
(C−CN), 23.7 (C−CH2). 3-Phenylpropanal Oxime (3q).21 White solid, 1.19 g, 80% yield.
2-Phenoxyacetonitrile (2k).6b Colorless oil, 320 mg, 80% yield. 1H 1
H NMR (300 MHz, CDCl3) δ 8.80 (br s, 1H, H−OH, major and
NMR (300 MHz, CDCl3): δ 7.40−7.33 (m, 2H, H-m-phenyl), 7.13− minor), 7.38 (t, J = 6.0 Hz, 1H, H−CHN, minor), 7.24−7.09 (m,
7.07 (m, 1H, H-p-phenyl), 7.02−6.97 (m, 2H, H-o-phenyl), 4.77 (s, 5H, H-phenyl, major and minor), 6.67 (t, J = 5.1 Hz, 1H, H−CH
2H, H−CH2). 13C NMR (75 MHz, CDCl3): δ 156.7 (C1-phenyl), N, major), 2.77−2.71 (m, 2H, H−C-3, major and minor), 2.67−2.59
130.0 (C-m-phenyl), 123.3 (C-p-phenyl), 115.1 (C-o-phenyl and C− (m, 2H, H−C-2, major), 2.48−2.41 (m, 2H, H−C-2, minor). 13C
CN), 53.7 (C−CH2). NMR (75 MHz, CDCl3) δ 151.8 (C−CHN, major), 151.5 (C−
Cyclohexanecarbonitrile (2l).17 Colorless oil, 285 mg, 87% yield. CHN, minor), 140.7 (C1-phenyl), 140.6 (C1-phenyl), 128.6,
1
H NMR (300 MHz, CDCl3): δ 2.61 (tt, J = 8.1, 3.9 Hz, 1H, H−C- 128.5, 128.4, 126.4 (C2−C6-phenyl), 32.9 (C-3, minor), 32.0 (C-3,
1), 1.89−1.80 (m, 2H, H−C-2 and H−C-6), 1.77−1.63 (m, 4H, H′− major), 31.3 (C-2, minor), 26.5 (C-2, major). The NMR peak
C-2 and H′−C-6, H−C-3 and H−C-5), 1.51−1.36 (m, 4H, H′−C-3 assignments were confirmed by the HMQC spectrum.


and H′−C-5, H−C-4). 13C NMR (75 MHz, CDCl3): δ 122.7 (C−
CN), 29.6 (C-2 and C-6), 28.1 (C-1), 25.3 (C-4), 24.2 (C-3 and C- ASSOCIATED CONTENT
5). The NMR peak assignments were confirmed by the HMQC
spectrum. *
S Supporting Information
2-Phenylbutanenitrile (2m).6b Yellow oil, 379 mg, 87% yield. 1H The Supporting Information is available free of charge on the
NMR (300 MHz, CDCl3): δ 7.42−7.29 (m, 5H, H-phenyl), 3.74 (t, J ACS Publications website at DOI: 10.1021/acs.joc.8b02190.
= 7.2 Hz, 1H, H−CHCN), 2.00−1.90 (m, 2H, H−CH2), 1.08 (t, J =
7.5 Hz, 3H, H−CH3). 13C NMR (75 MHz, CDCl3): δ 135.9 (C1- 1
H and 13C NMR spectra of all the synthesized
phenyl), 129.1 (C-m-phenyl), 128.1 (C-p-phenyl), 127.4 (C-o- compounds (PDF)
phenyl), 120.9 (C−CN), 39.0 (H-CHCN), 29.3 (C−CH2), 11.6


(C−CH3). The NMR peak assignments were confirmed by the
HMBC spectrum. AUTHOR INFORMATION
Adamantane-1-carbonitrile (2n).15 White solid, 396 mg, 82%
yield. 1H NMR (300 MHz, CDCl3): δ 2.04 (br, 9H, H−CH2−β-CN Corresponding Authors
and H−CH-γ-CN), 1.77−1.68 (m, 6H, H−CH2−δ-CN). 13C NMR *Tel and Fax: +8610 68984545. E-mail: tianhy@btbu.edu.cn.
(75 MHz, CDCl3): δ 125.4 (C−CN), 40.0 (C-β-CN), 35.9 (C-δ- *Tel and Fax: +8610 68984545. E-mail: sunbg@btbu.edu.cn.
CN), 30.3 (C-α-CN), 27.2 (C-γ-CN). The NMR peak assignments
were confirmed by the dept135 and HMQC spectra. ORCID
(S)-tert-Butyl 1-Cyano-2-methylpropylcarbamate (2o).18 White Hongyu Tian: 0000-0002-7117-6420
solid, 493 mg, 83% yield. 1H NMR (300 MHz, CDCl3, 40 °C) δ 4.93 Baoguo Sun: 0000-0003-4326-8237
(br d, J = 8.7 Hz, 1H, H-NH), 4.42 (br s, 1H, H−CHCN), 2.01
(octet, J = 6.6 Hz, 1H, H−CH(CH3)2), 1.46 (s, 9H, H−OC(CH3)3), Notes
1.08 (d, J = 6.6 Hz, 3H, H−CH(CH3)2), 1.06 (d, J = 6.3 Hz, 3H, H− The authors declare no competing financial interest.
E DOI: 10.1021/acs.joc.8b02190
J. Org. Chem. XXXX, XXX, XXX−XXX
The Journal of Organic Chemistry Note

■ ACKNOWLEDGMENTS
Financial support from the National Key Research and
oximes. Tetrahedron 2012, 68, 2899−2905. (b) Shipilovskikh, S. A.;
Vaganov, V. Y.; Denisova, E. I.; Rubtsov, A. E.; Malkov, A. V.
Dehydration of amides to nitriles under conditions of a catalytic
Development Program (2016YFD0400801), the Beijing Appel reaction. Org. Lett. 2018, 20, 728−731. (c) Liu, R. Y.; Bae, M.;
Postdoctoral Research Foundation (2017-22-011), and the Buchwald, S. L. Mechanistic insight facilitates discovery of a mild and
Importation and Development of High-Caliber Talents Project efficient copper-catalyzed dehydration of primary amides to nitriles
of Beijing Municipal Institutions (CIT&TCD20140306) is using hydrosilanes. J. Am. Chem. Soc. 2018, 140, 1627−1631.
gratefully acknowledged. (7) Mancuso, A. J.; Huang, S.; Swern, D. Oxidation of long-chain


and related alcohols to carbonyls by dimethyl sulfoxide ″activated″ by
REFERENCES oxalyl chloride. J. Org. Chem. 1978, 43, 2480−2482.
(8) Jadhav, S. B.; Ghosh, U. A simple, rapid, and efficient protocol
(1) (a) Smith, M. B.; March, J. Eliminations. In March’s Advanced for the synthesis of methylthiomethyl esters under Swern oxidation
Organic Chemistry: Reactions, Mechanisms and Structure, 7th ed.; John conditions. Tetrahedron Lett. 2007, 48, 2485−2487.
Wiley & Sons: Hoboken, NJ, 2013. (b) Kukushkin, V. Y.; Pombeiro, (9) Nakajima, N.; Ubukata, M. Preparation of nitriles from primary
A. J. L. Additions to metal-activated organonitriles. Chem. Rev. 2002, amides under Swern oxidation conditions. Tetrahedron Lett. 1997, 38,
102, 1771−1802. (c) Gaspar, B.; Carreira, E. M. Mild cobalt-catalyzed 2099−2102.
hydrocyanation of olefins with tosyl cyanide. Angew. Chem., Int. Ed. (10) Zhang, T.; Dai, Y.; Cheng, S.; Liu, Y.; Yang, S.; Sun, B.; Tian,
2007, 46, 4519−4522. (d) Hummel, J. R.; Boerth, J. A.; Ellman, J. A. H. A facile method for the sulfenyllactonization of alkenoic acids
Transition-metal-catalyzed C-H bond addition to carbonyls, imines, using dimethyl sulfoxide activated by oxalyl chloride. Synthesis 2017,
and related polarized π bonds. Chem. Rev. 2017, 117, 9163−9227. 49, 1380−1386.
(e) Pearson-Long, M. S. M.; Boeda, F.; Bertus, P. Double addition of (11) Ding, R.; Lan, L.; Li, S.; Liu, Y.; Yang, S.; Tian, H.; Sun, B. A
organometallics to nitriles: toward an access to tertiary carbinamines. novel method for the chlorolactonization of alkenoic acids using
Adv. Synth. Catal. 2017, 359, 179−201. (f) Lindsay-Scott, P. J.; diphenyl sulfoxide/oxalyl chloride. Synthesis 2018, 50, 2555−2566.
Gallagher, P. T. Synthesis of heterocycles from arylacetonitriles: (12) Ding, R.; Li, J.; Jiao, W.; Han, M.; Liu, Y.; Tian, H.; Sun, B. A
powerful tools for medicinal chemists. Tetrahedron Lett. 2017, 58, highly efficient method for bromination of alkenes, alkynes, and
2629−2635. ketones using dimethyl sulfoxide and oxalyl bromide. Synthesis 2018,
(2) (a) Fleming, F. F.; Yao, L.; Ravikumar, P. C.; Funk, L.; Shook, B. DOI: 10.1055/s-0037-1609560.
C. Nitrile-containing pharmaceuticals: efficacious roles of the nitrile (13) Ding, R.; Li, Y.; Liu, Y.; Sun, B.; Yang, S.; Tian, H. Synthesis of
pharmacophore. J. Med. Chem. 2010, 53, 7902−7917. (b) Song, H. Y.; butenolides by reactions of 3-alkenoic acids with diphenyl sulfoxide/
Yang, J. Y.; Suh, J. W.; Lee, H. S. Acaricidal activities of apiol and its oxalyl chloride. Flavour Fragrance J. 2018, DOI: 10.1002/ffj.3464.
derivatives from petroselinum sativum seeds against dermatopha- (14) (a) Zhang, X.; Waymouth, R. M. Zwitterionic ring opening
goides pteronyssinus, dermatophagoides farinae, and tyrophagus polymerization with isothioureas. ACS Macro Lett. 2014, 3, 1024−
putrescentiae. J. Agric. Food Chem. 2011, 59, 7759−7764. 1028. (b) Carafa, M.; Mesto, E.; Quaranta, E. DBU-promoted
(c) Kleemann, A.; Engel, J.; Kutscher, B.; Reichert, D. Pharmaceutical nucleophilic activation of carbonic acid diesters. Eur. J. Org. Chem.
Substances: Syntheses, Patents and Applications, 5th ed.; Thieme: 2011, 2458−2465.
Stuttgart, 2008. (d) Gouda, M. A.; Hussein, B. H. M.; Helal, M. H.; (15) Chiba, S.; Zhang, L.; Ang, G. Y.; Hui, B. W. Generation of
Salem, M. A. A review: synthesis and medicinal importance of iminyl copper species from α-azido carbonyl compounds and their
nicotinonitriles and their analogs. J. Heterocyclic Chem. 2018, 55, catalytic C-C bond cleavage under an oxygen atmosphere. Org. Lett.
1524−1553. (e) David, J. R. Chemistry and Technology of Flavors and 2010, 12, 2052−2055.
Fragrances, 1st ed.; Blackwell: Oxford, 2005. (16) Campbell, J. A.; McDougald, G.; McNab, H.; Rees, L. V. C.;
(3) (a) Najam, T.; Shah, S. S. A.; Mehmood, K.; Din, A. U.; Rizwan, Tyas, R. G. Laboratory-scale synthesis of nitriles by catalyzed
S.; Ashfaq, M.; Shaheen, S.; Waseem, A. An overview on the progress dehydration of amides and oximes under flash vacuum pyrolysis
and development on metals/non-metal catalyzed cyanation reactions. (FVP) conditions. Synthesis 2007, 3179−3184.
Inorg. Chim. Acta 2018, 469, 408−423. (b) Nonn, M.; Remete, A. M.; (17) Li, Y.; Liao, B.; Chen, H.; Liu, S. Ligand-free nickel-catalyzed
Fülöp, F.; Kiss, L. Recent advances in the transformations of conversion of aldoximes into nitriles. Synthesis 2011, 2639−2643.
cycloalkane-fused oxiranes and aziridines. Tetrahedron 2017, 73, (18) Mangette, J. E.; Johnson, M. R.; Le, V.; Shenoy, R. A.; Roark,
5461−5483. (c) Kouznetsov, V. V.; Galvis, C. E. P. Strecker reaction H.; Stier, M.; Belliotti, T.; Capiris, T.; Guzzo, P. R. The preparation of
and α-amino nitriles: recent advances in their chemistry, synthesis, optically active α-amino 4H-[1,2,4]oxadiazol-5-ones from optically
and biological properties. Tetrahedron 2018, 74, 773−810. (d) Amal active α-amino acids. Tetrahedron 2009, 65, 9536−9541.
Joseph, P. J.; Priyadarshini, S. Copper-mediated C-X functionalization (19) Hoang, C. T.; Bouillere, F.; Johannesen, S.; Zulauf, A.; Panel,
of aryl halides. Org. Process Res. Dev. 2017, 21, 1889−1924. (e) Yan, C.; Pouilhes, A.; Gori, D.; Alezra, V.; Kouklovsky, C. Amino acid
G.; Zhang, Y.; Wang, J. Recent advances in the synthesis of aryl nitrile homologation by the Blaise reaction: a new entry into nitrogen
compounds. Adv. Synth. Catal. 2017, 359, 4068−4105. heterocycles. J. Org. Chem. 2009, 74, 4177−4187.
(4) (a) Song, X.; Qiu, Y.; Liu, X.; Liang, Y. Recent advances in the (20) Mori, N.; Togo, H. Direct oxidative conversion of primary
tandem reaction of azides with alkynes or alkynols. Org. Biomol. Chem. alcohols to nitriles using molecular iodine in ammonia water. Synlett
2016, 14, 11317−11331. (b) Kolle, S.; Batra, S. Transformations of 2005, 1456−1458.
alkynes to carboxylic acids and their derivatives via CC bond (21) Minakata, S.; Okumura, S.; Nagamachi, T.; Takeda, Y.
cleavage. Org. Biomol. Chem. 2016, 14, 11048−11060. Generation of nitrile oxides from oximes using t-BuOI and their
(5) (a) Yadav, L. D. S.; Srivastava, V. P.; Patel, R. Bromodime- cycloaddition. Org. Lett. 2011, 13, 2966−2969.
thylsulfonium bromide (BDMS): a useful reagent for conversion of
aldoximes and primary amides to nitriles. Tetrahedron Lett. 2009, 50,
5532−5535. (b) Sharghi, H.; Sarvari, M. H. Graphite as an efficient
catalyst for one-step conversion of aldehydes into nitriles in dry
media. Synthesis 2003, 243−246. (c) Singh, M. K.; Lakshman, M. K. A
simple synthesis of nitriles from aldoximes. J. Org. Chem. 2009, 74,
3079−3084. (d) Gucma, M.; Golebiewski, W. M. Convenient
conversion of aldoximes into nitriles with N-chlorosuccinimide and
pyridine. Synthesis 2008, 1997−1999.
(6) (a) Denton, R. M.; An, J.; Lindovska, P.; Lewis, W.
Phosphonium salt-catalysed synthesis of nitriles from in situ activated

F DOI: 10.1021/acs.joc.8b02190
J. Org. Chem. XXXX, XXX, XXX−XXX

You might also like