Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Research article

Received: 01 December 2015, Revised: 12 January 2016, Accepted: 18 January 2016, Published online in Wiley Online Library: 18 February 2016

(wileyonlinelibrary.com) DOI: 10.1002/poc.3543

A DFT study of reduction of nitrobenzene to


aniline with SnCl2 and hydrochloric acid
Shinichi Yamabea* and Shoko Yamazakib

A fundamental reduction reaction, nitrobenzene to aniline in SnCl2 and hydrochloric acid, was investigated by density

functional theory (DFT) calculations. First, the change of SnCl2 → SnCl2
4 → Cl4SnH was discussed, and the reaction path of

SnCl4 + H3O → Cl4SnH + H2O was obtained. Starting from nitrobenzene, six elementary processes were found so as to
2 +

arrive at the protonated aniline. The hydride ion from Cl4SnH is connected always to the cationic nitrogen, and the proton
is always to oxygens. An intermediate Ph–N+H2OH was obtained, which is isomerized to the para O–H adduct protonated
imine via the Bamberger rearrangement. This species may undergo the H acceptance at the sp2 N+H2 center. In the nitroben-
zene reduction, the proton enhances the electrophilicity of the nitrogen center, which makes the hydride shift ready. N–H
bonds are formed, and N–O bonds are cleaved both by the proton attach and subsequent H2O elimination and by the formal
[1,5] OH shift. Copyright © 2016 John Wiley & Sons, Ltd.

Keywords: density functional theory calculations; nitrobenzene; reduction; transition state; water cluster

INTRODUCTION
The reduction of nitrobenzene to aniline is a well-known reaction
and was first performed by Zinin in 1842 using inorganic sulfide
as a reductant (Zinin reaction).[1a] The Bechamp reduction used
iron and hydrochloric acid.[1b] Nowadays, the reaction is con-
ducted usually in the condition of Eqn (1).
Scheme 2. A reduction mechanism thought so far

undec-7-ene,[10] Pd(OAc)2/polymethylhydrosiloxane,[11] Sm/I2,[12]


Sm/1,1’-dioctyl-4,4’-bipyridinium dibromide,[13] Sm/NH4Cl,[14] Cu
nanoparticles/HCOONH4,[15] S8/NaHCO3,[16] HI,[17] and silane/oxo-
ruthenium complexes[18] have been used for the reduction.
The stannous chloride (SnCl2) was also used as shown in In spite of those extensive experimental studies, the precise
Scheme 1.[2] mechanism of the fundamental reaction, nitrobenzene → aniline,
The reduction has been presumed to proceed with two interme- has been still veiled. In particular, it is still an open question how
diates, nitrosobenzene and phenyl hydroxylamine (Scheme 2).[3] the two N–O bonds are converted to the N–H ones in Scheme 2.
Many methods have been developed to carry out the selec- In order to solve the naive and important question, density func-
tive reduction of aromatic nitro compounds. They include cata- tional theory (DFT) calculations were performed by the use of a
lytic hydrogenation,[4] sodium trimethylsilanethiolate,[5] sodium model composed of nitrobenzene, SnCl2, H3O+, and the water
borohydride/catalyst,[6] hydrazine/catalyst,[7] formates/catalyst,[8] cluster. The model follows the condition in Scheme 1(b), that is,
and metals such as iron and zinc besides tin in Eqn (1).[9] Moreover, SnCl2 in the concentrated. HCl aqueous solution. The metal tin
a variety of other chemical systems, Mo(CO)6/1,8-diazabicyclo[5.4.0] in Eqn (1), which is a standard reductant, was not adopted in
the present DFT calculations, because the aggregate solid state
of tin is difficult to express in the model construction.

* Correspondence to: Shinichi Yamabe, Graduate School of Materials Science,


Nara Institute of Science and Technology (NAIST), Takayama, Ikoma, Nara
630-0192, Japan.
E-mail: yamabe@ms.naist.jp

a S. Yamabe
Graduate School of Materials Science, Nara Institute of Science and Technol-
ogy (NAIST), Takayama, Ikoma, Nara 630-0192, Japan

b S. Yamazaki
361

Scheme 1. Two reported reactions using the SnCl2 reducing agent in Department of Chemistry, Nara University of Education, Takabatake-cho, Nara
Refs. 2(a) and (b) 630-8528, Japan

J. Phys. Org. Chem. 2016, 29 361–367 Copyright © 2016 John Wiley & Sons, Ltd.
S. YAMABE AND S. YAMAZAKI

effect (solvent = water). Transition states (TSs) were sought first


by partial optimizations at bond-interchange regions. Second,
by the use of Hessian matrices, TS geometries were optimized.
They were characterized by vibrational analyses, which checked
whether the obtained geometries have single imaginary
frequencies (ν‡s). From TSs, reaction paths were traced by the
intrinsic reaction coordinate method[25,26] to obtain the energy-
minimum geometries. Energy changes (ΔEs) along elementary
processes were shown by the use of the sum of electronic and
zero-point vibrational energies, “Et + ZPE”. All the calculations
were carried out using the Gaussian 09 program package.[27]
The computations were performed at the Research Center for
Scheme 3. The Sn(II) → Sn(IV) oxidation process along with the coordi-
 Computational Science, Okazaki, Japan.
nation of two chloride ions and the formation of a hydride donor HSnCl4

METHOD OF CALCULATIONS A MODEL OF THE SN(II) → SN(IV) OXIDATION


PROCESS
The reacting systems were investigated by DFT calculations. The
B3LYP[19,20] method was used. The basis sets employed were Stannous chloride is isolobal with dichlorocarbene CCl2. Therefore,
6-31++G(d) for C, H, O, N, and Cl atoms. For the Sn atom, the the lowest unoccupied molecular orbital of SnCl2 is composed
relativistic core potential of SDD basis set[21] was used. Geometry mainly of 5p(π) of tin as shown in Figure S1. In addition, owing
optimizations and subsequent vibrational analyses were carried to the large difference of electronegativities of Sn(1.8) and Cl
out including the polarizable continuum model[22–24] solvent (3.0), tin is very cationic (charge = +1.17 by natural population

2 +  + 2- +
Figure 1. A proton shift path to form a Cl4Sn–H species in SnCl4 + (H3O )2(H2O)6 → Cl4SnH + H3O (H2O)7. The precursor of the SnCl4 + (H3O )2
+ (H2O)6 is SnCl2 + 2HCl + (H2O)8 (before SnCl2 is put into the hydrochloric-acid solution). All the geometries of precursors, TSs and products are shown
in Supporting Information. Bond-interchange regions are marked by orange color
362

+
Scheme 4. Models of the hydride transfer to Ph–NO2 enhanced by the proton transfer from one H3O with the other one (a) and two (b) hydronium
catalysts in Fig. 2. Bold empty arrows indicate proton and hydride transfers

wileyonlinelibrary.com/journal/poc Copyright © 2016 John Wiley & Sons, Ltd. J. Phys. Org. Chem. 2016, 29 361–367
A DFT STUDY OF REDUCTION OF NITROBENZENE TO ANILINE

bilization of ΔG°, 16.74 kcal/mol, was obtained. Thus, formation of


the SnCl24 is a likely process. This di-anion is not of the tetrahedral
geometry but of the one shown in Scheme 3 (with the axial linear
Cl–Sn–Cl bond). The species SnCl2 4 has a highly nucleophilic 5s
lone-pair orbital (Figure S2) to which a proton is captured readily.
A Cl4Sn–H species is formed. Next, if a hydride ion is evolved from
Cl4SnH, SnCl4 is afforded. The step of the proton capture, that is,

4 + H → Cl4Sn–H , in the present hypothesis was examined.
SnCl2 +

A model of SnCl4 + (H3O+)2(H2O)6 → Cl4SnH + H3O+(H2O)7 was


2

employed. The proton transfer path is shown in Fig. 1.


The activation energy, ΔE‡ = +16.7 kcal/mol, is small, and the left-
ward proton shift is a ready process. On the other hand, the reaction
energy, ΔE = +6.98 kcal/mol, is somewhat endothermic, which how-
ever may be overcome by the exothermicity of subsequent hydride-
shift processes. Before the shift (the left side of Fig. 1), the SnCl4
moiety has the natural charge 1.998, and the SnCl2 4 character
was confirmed. Similarly, after the shift, the Cl4SnH moiety has the
charge 0.944, and the Cl4SnH one was confirmed.
The reactant in Fig. 1 is re-written as SnCl2 + 2HCl + (H2O)8 in
the neutral form. Another reactant model composed of SnCl2
+ HCl + (H2O)7, that is, SnCl +
3 + H3O (H2O)6, was examined for
the hydride transfer. Figure S6 presents the result. The activation
and reaction energies are +24.72 and +22.21 kcal/mol, respec-
tively. They are larger than those in Fig. 1, and accordingly, the
model of SnCl2 + HCl + (H2O)7 is unlikely. Thus, the ion pair
[(H3O+)Cl]2 toward SnCl2 is needed to cause the ready hydride
shift. In the next section, it will be investigated how the nucleo-
phile Cl4SnH reacts with nitrobenzene involving potential inter-
mediates in Scheme 2 under the acidic condition.

Figure 2. Transition states of the hydride (H1) shift to the nitro nitrogen
ELEMENTARY PROCESSES FROM NITROBEN-
atom assisted by the proton (H2) shift onto one oxygen atom (O2) and ZENE TO PROTONATED ANILINE

the proton catalyst on the other oxygen atom. In (a), Cl4SnH + Ph–NO2
+ +  + Scheme 4 exhibits two reaction models for the hydride transfer
+ (H3O )2 + (H2O)7 → Cl4Sn + Ph–N (OH)H–O + H3O + (H2O)8. In (b),
 + + 
Cl4SnH + Ph–NO2 + (H3O )3 + (H2O)9 → Cl4Sn + Ph–N (OH)H–O + (H3O )
+
from Cl4SnH. To promote the nucleophilic attack, the nitrogen
2 + (H2O)10. A reaction formula representing schematically the central part is atom needs to be electrophilic as much as possible. By the coor-
shown below two TS geometries. In Scheme 4, reaction models of (a) and dination of H3O+ to lone-pair electrons of oxygen of the nitro
(b) are pictorially shown group, the electrophilicity is enhanced. In (a), two H3O+ are coor-
dinated, one of which gives the proton to one oxygen atom. To
analysis). The chloride ion in hydrochloric acid may be bound to free O–H bonds of the two H3O+, five water molecules are hy-
the lowest unoccupied molecular orbital of SnCl2 (Scheme 3). drogen bonded for stabilization of the system. The other lone-
Changes of Gibbs and entropy changes in the process, SnCl2 pair electrons of the nitro-group oxygen are linked to two water
+ 2Cl → SnCl2 4 , were calculated at the condition T = 298.15 K and molecules, respectively. In (b), three H3O+ were adopted to in-
P = 1 atm. In spite of large entropy loss, 42.32 cal/(mol·K), large sta- crease further the electrophilicity of the nitro-group nitrogen.

363

+  + +
Figure 3. A proton shift transition state in SnCl4 + Ph–N (OH)H–O + H3O + (H2O)8 → SnCl4 + Ph–N (OH)2H + (H2O)9

J. Phys. Org. Chem. 2016, 29 361–367 Copyright © 2016 John Wiley & Sons, Ltd. wileyonlinelibrary.com/journal/poc
S. YAMABE AND S. YAMAZAKI

+  + +
Figure 4. A proton shift path to form the N-protonated nitrosobenzene, Ph-N H(OH)–O + SnCl4 + H3O + (H2O)8 → Ph–N (¼O)H + SnCl4 + (H2O)10

+  +
Figure 5. The conversion path from N-protonated nitrosobenzene to N-protonated phenylhydroxylamine, Ph–N (¼O)H + Cl4SnH + H3O + (H2O)9 → Ph–
+
N (OH)H2 + Cl4Sn + (H2O)10

Scheme 5. The Bamberger rearrangement. (a) The traditionally accepted mechanism.[29] (b) The results of our recent computational study[30]

It should be noted that there are many other initial geometries Figure 2 shows two TSs of the hydride transfer, Sn....H1....N and the
for the hydride transfer. For instance, one H3O+ may be situated concomitant proton one, O3.....H2....O2. In Fig. 2(a), the nitro-group
in the form, Ph–NO–O....H2O…H3O+. Therefore, models shown oxygen O1 is coordinated by one H3O+, which is the acid catalyst
in Schemes 4(a) and (b) are not unique. However, they are to enhance the electrophilicity of N.
364

thought to be likely to represent the simultaneous movement In Fig. 2(b), O1 is by two hydronium ions. Activation energies
of the proton and hydride. are +35.33 kcal/mol (a) and +33.76 kcal/mol (b), respectively.

wileyonlinelibrary.com/journal/poc Copyright © 2016 John Wiley & Sons, Ltd. J. Phys. Org. Chem. 2016, 29 361–367
A DFT STUDY OF REDUCTION OF NITROBENZENE TO ANILINE

These relatively large energies would correspond to the heat The proton shift leads to Ph–N+(¼O)H species and H2O with a
(boiling in Ref. 2b) condition. After each TS, an intermediate large exothermic energy, 19.54 kcal/mol. The species, N-
Ph–N+H(OH)–O is formed. This species is thought to be primar- protonated nitrosobenzene, may be subject to the hydride shift
ily subject to the protonation, Ph–N+H(OH)–O + H+ → Ph–N+H from Cl4SnH (Fig. 5), which has the pattern similar to that of
(OH)2. Its TS is shown in Fig. 3. Fig. 2.
Activation and reaction energies are very small (1.40 The hydride shift takes place along with the proton one,
and 0.52 kcal/mol, respectively), and the protonated interme- yielding N-protonated phenylhydroxylamine, Ph–N+(OH)H2. This
diate Ph–N+H(OH)2 stays merely in equilibrium with Ph–N+H cation is known well in the Bamberger rearrangement.[28] By
(OH)–O. Alternatively, if the proton is shifted to the hydroxyl the rearrangement, N-phenylhydroxylamine is converted to
group of Ph–N+H(OH)–O, a TS geometry shown in Fig. 4 was para-aminophenol in the aqueous sulfuric acid. The mechanism
obtained. including a nitrenium intermediate is shown in Scheme 5(a),


365

+ + + +
Figure 6. Three steps starting from Ph–N (OH)H2 to the product Ph–NH3 . Ph–N (OH)H2 + H3O + (H2O)9 + Cl4SnH → TS(a) → para-OH added
+ +   + + +
Ph = N H2 + H3O + (H2O)9 + Cl4SnH → TS(b) → para-OH added Ph –NH3 [zwitterion] + H3O + (H2O)9 + SnCl4 → TS(c) → Ph–NH3 + (H2O)11 + SnCl4

J. Phys. Org. Chem. 2016, 29 361–367 Copyright © 2016 John Wiley & Sons, Ltd. wileyonlinelibrary.com/journal/poc
S. YAMABE AND S. YAMAZAKI

bond cleavage and C–O bond formation occur simultaneously


by aid of proton transfers. If the resultant cation HO(H)–
C6H4¼N+H2 underwent the deprotonation, the Bamberger-
rearrangement product para-amino-phenol HO–C6H4–NH2
would be yielded. However, in the concentrated H–Cl solution,
the reaction, that is, the proton dispersal from the cation would
be unfavorable. Alternatively, the protonated imine N center
C¼N+H2 of the cation may be subject to the hydride shift, of
which TS is shown in Fig. 6(b). In fact, the activation energy of
TS(b) in Fig. 6 is smaller than that of the deprotonation shown
in Figure S5 (in Supporting Information). The NH+3 moiety is
completed in the zwitterion intermediate after TS(b). Finally, a
proton shift to the para OH group attached to the anionic
six-membered ring leads to the product, protonated aniline,
and H2O (Fig. 6(c)).
An important result found in the present calculations is that
the hydride shift occurs toward the π* direction of the N+
center of the substrate, which is shown commonly in Figs. 2,
5, and 6(b). The SN2 type path at the tetrahedral center shown
in Scheme 6 could not be found. On the other hand, an SN1
Scheme 6. An SN2 path concomitant with the front side proton transfer. path involving the aniline di-cation intermediate was obtained
The path could not be obtained. Bold arrows show the movement of H and is shown in Figure S4 (Supporting Information). However,
and O atoms the SN1 path was found to be less favorable energetically than
TS(a) in Fig. 6.
which was suggested by Ingold et al.[29] However, it was not In the Bamberger rearrangement, ortho-chloro-amino and
solved why ortho-aminophenol is not afforded by the rear- para-chloro-amino derivatives are also yielded in hydrochloric
rangement. Recently, the long-term mystery of the selectivity acid.[28b] In our previous work,[30] TS geometries of the Ph–N+H2
was investigated by our computational study.[30] Its result (OH) + H3O+(H2O)14 + Cl → para- (and ortho-) Cl–C6H5¼N+H2
revealed that the size of the hydrogen-bond network for the + (H2O)16 were determined. In the conc. In HCl solution, the
proton transfer starting from Ph–N+H2(OH) is fit for the para Cl-adduct intermediate would be formed along with para-HO–C-
6H5¼N H2 by the rearrangement in Fig. 6(a). The subsequent
+
product formation (Scheme 5(b)).
The Bamberger-rearrangement path is shown in Fig. 6(a). N–O step from the intermediate was sought. Figure S3 shows a TS
366

Scheme 7. Reduction of nitrobenzene to protonated aniline via supply of proton and hydride ion and water elimination. Numbers with two decimal
places are energy changes in kcal/mol

wileyonlinelibrary.com/journal/poc Copyright © 2016 John Wiley & Sons, Ltd. J. Phys. Org. Chem. 2016, 29 361–367
A DFT STUDY OF REDUCTION OF NITROBENZENE TO ANILINE

geometry where the hydride shift and the Cl elimination from H. Junge, R. Llusar, M. Beller, J. Am. Chem. Soc. 2011, 133,
the intermediate takes place at the same time. Thus, the reac- 12875–12879.
[9] (a) A. B. Gamble, J. Garner, C. P. Gordon, S. M. J. O’Conner, P. A.
tion from Ph–N+H2(OH) to Ph–N+H3 proceeds via HO and Cl- Keller, Synth. Commun. 2007, 37, 2777–2786; (b) Y. Liu, Y. Lu, M.
adduct C6H5¼N+H2 intermediates (the benzene aromatic ring Prashad, O. Repic, T. J. Blacklock, Adv. Synth. Catal. 2005, 347,
destroyed transiently). 217–219; (c) H. Mahdavi, B. Tamami, Synth. Commun. 2005, 35,
1121–1127; (d) P. De, Synlett. 2004, 1835–1837; (e) L. Wang, P. Li,
Z. Wu, J. Yan, M. Wang, Y. Ding, Synthesis. 2003, 2001–2004; (f) F.
CONCLUDING REMARKS A. Kahn, J. Dash, C. Sudheer, R. K. Gupta, Tetrahedron Lett. 2003,
44, 7783–7787; (g) S. M. Kelly, B. H. Lipshutz, Org. Lett. 2014, 16,
98–101.
In this study, a fundamental reduction, nitrobenzene to aniline in [10] J. Spencer, R. P. Rathnam, H. Patel, N. Anjum, Tetrahedron. 2008, 64,
the SnCl2, and hydrochloric acid, was investigated by DFT calcu- 10195–10200.

lations. In Scheme 3, the change of SnCl2 → SnCl2 4 → Cl4SnH [11] R. J. Rahaim, R. E. Maleczka, Org. Lett. 2005, 7, 5087–5090.
was assumed, and it was explained that the oxidized species [12] B. K. Banik, C. Mukhopadhyay, M. S. Venkatraman, F. F. Becker,
Tetrahedron Lett. 1998, 39, 7243–7246.
SnCl4 is formed by evolving a hydride ion. The path of
 [13] C. Yu, B. Liu, L. Hu, J. Org. Chem. 2001, 66, 919–924.
4 + H3O → Cl4SnH + H2O was obtained. A summary of
SnCl2 +
[14] M. K. Basu, F. F. Becker, B. K. Banik, Tetrahedron Lett. 2000, 41,
calculations is illustrated in Scheme 7, where H comes from 5603–5606.
Cl4Sn–H and H+ comes from H3O+. [15] A. Saha, B. Ranu, J. Org. Chem. 2008, 73, 6867–6870.
Six elementary processes (except that in Fig. 3) were found. [16] M. A. McLaughlin, D. M. Barnes, Tetrahedron Lett. 2006, 47,
9095–9097.
The hydride ion is connected always to the cationic nitrogen, [17] J. S. Dileep, M. M. Ho, T. Toyokuni, Tetrahedron Lett. 2001, 42,
and the proton is always to oxygens. A striking result is at the 5601–5603.
conversion of Fig. 6(a), where Ph–N+H2OH is isomerized to the [18] R. G. Noronha, C. C. Romao, A. C. Fernandes, J. Org. Chem. 2009, 74,
para O–H adduct protonated imine via the Bamberger rear- 6960–6964.
rangement. This cation may undergo the H acceptance at the [19] A. D. J. Becke, Chem. Phys. 1993, 98, 5648–5652.
[20] C. Lee, W. Yang, R. G. Parr, Phys. Rev. B. 1988, 37, 785–789.
sp2 N+H2 center. [21] A. Bergner, M. Dolg, W. Kuchle, H. Stoll, H. Preuss, Mol. Phys. 1993,
This work showed that the nitrobenzene reduction occurs by the 80, 1431–1441.
cooperation of H and H+ movement. The proton enhances the [22] E. Cances, B. Mennucci, J. Tomasi, J. Chem. Phys. 1997, 107,
electrophilicity of the nitrogen center, which makes the hydride shift 3032–3041.
[23] M. Cossi, V. Barone, B. Mennucci, J. Tomasi, Chem. Phys. Lett. 1998,
ready. Changing nitrosobenzene and phenylhydroxylamine into re- 286, 253–260.
spective protonated species in Scheme 2 would give a plausible [24] B. Mennucci, J. Tomasi, J. Chem. Phys. 1997, 106, 5151–5158.
mechanism. [25] K. Fukui, J. Phys. Chem. 1970, 74, 4161–4163.
[26] C. Gonzalez, H. B. Schlegel, J. Chem. Phys. 1989, 90, 2154–2161.
[27] M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb,
REFERENCES J. R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A.
Petersson, H. Nakatsuji, M. Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov,
[1] (a) N. Zinin, J. Prakt. Chem. 1842, 27, 140–153; (b) A. Bechamp, Ann. J. Bloino, G. Zheng, J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota,
Chim. Phys. 1854, 42, 186–196. R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao,
[2] (a) L. I. Smith, Org. Synth. Coll. 1943, 2, 254; (b) J. S. Buck, W. S. Ide, H. Nakai, T. Vreven, J. A. Montgomery Jr. , J. E. Peralta, F. Ogliaro,
Org. Synth. Coll. 1943, 2, 130. M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov,
[3] (a) F. Z. Harber, Elektrochem. 1898, 4, 506–513; (b) J. March, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J. C. Burant,
Advanced Organic ChemistryWiley-Interscience, New York, 1992, S. S. Iyengar, J. Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene,
1216–1217. J. E. Knox, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts,
[4] (a) M. Takasaki, Y. Motoyama, K. Higashi, S.-H. Yoon, I. Mochida, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli,
H. Nagashima, Org. Lett. 2008, 10, 1601–1604; (b) M. L. Kantam, J. W. Ochterski, R. L. Martin, K. Morokuma, V. G. Zakrzewski,
T. Bandyopadhyay, A. Rahman, J. Mol. Catal. A: Chem. 1998, 133, G. A. Voth, P. Salvador, J. J. Dannenberg, S. Dapprich, A. D. Daniels,
293–295; (c) Z. Wei, J. Wang, S. Mao, D. Su, H. Jin, Y. Wang, F. Xu, Ö. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski, D. J. Fox, Gaussian
H. Li, Y. Wang, ACS Catal. 2015, 5, 4783–4789. 09, Revision D.01, Gaussian, Inc., Wallingford CT, 2009.
[5] J. R. Hwu, F. F. Wong, M. J. Shiao, J. Org. Chem. 1992, 57, 5254–5255. [28] (a) E. Bamberger, Chem. Ber. 1894, 27, 1347–1350; (b) E. Bamberger,
[6] (a) P. D. Ren, S. F. Pan, T. W. Dong, S. H. Wu, Synth. Commun. 1995, Chem. Ber. 1894, 27, 1548–1557.
25, 3799–3803; (b) K. P. Chary, S. R. Ram, D. S. Iyengar, Synlett. 2000, [29] H. E. Heller, E. D. Hughes, C. K. Ingold, Nature. 1951, 168, 909–910.
683–685; (c) B. Zeynizadeh, D. Setamdideh, Synth. Commun. 2006, [30] S. Yamabe, G. Zeng, W. Guan, S. Sakaki, Beilstein J. Org. Chem. 2013,
36, 2699–2704. 9, 1073–1082.
[7] (a) S. P. Kumbhar, J. Sanchez-Valente, F. Figueras, Tetrahedron Lett.
1998, 39, 2573–2574; (b) A. Vass, J. Dudas, J. Toth, R. S. Varma,
Tetrahedron Lett. 2001, 42, 5347–5349; (c) Q. Shi, R. Lu, K. Jin,
Z. Zhang, D. Zhao, Green Chem. 2006, 8, 868–870; (d) D. Cantillo,
M. M. Moghaddam, C. O. Kappe, J. Org. Chem. 2013, 78, 4530–4542;
SUPPORTING INFORMATION
(e) C. Jiang, Z. Shang, X. Liang, ACS Catal. 2015, 5, 4814–4818.
[8] a A. B. Taleb, G. Jenner, J. Organomet. Chem. 1993, 456, 263–269; (b) Additional supporting information can be found in the online
G. Wienhöfer, I. Sorribes, A. Boddien, F. Westerhaus, K. Junge, version of this article at the publisher’s website.
367

J. Phys. Org. Chem. 2016, 29 361–367 Copyright © 2016 John Wiley & Sons, Ltd. wileyonlinelibrary.com/journal/poc

You might also like