You are on page 1of 12

INORGANIC CHEMISTRY

FRONTIERS
View Article Online
RESEARCH ARTICLE View Journal | View Issue

Cobalt complexes of pyrrolecarboxamide ligands


Published on 09 December 2016. Downloaded by Universitat de Barcelona on 8/27/2020 5:10:40 PM.

Cite this: Inorg. Chem. Front., 2017,


as catalysts in nitro reduction reactions: influence
4, 324 of electronic substituents on catalysis and
mechanistic insights†
Sunil Yadav, Sushil Kumar and Rajeev Gupta*

This work presents two square-planar Co(III) complexes within a N4 coordination sphere created by two
related pyrrolecarboxamide ligands. Both complexes functioned as catalysts for the hydrazine-mediated
reduction of aromatic, heterocyclic, as well as aliphatic nitro substrates. Binding studies provided evidence
Received 20th September 2016, that the square-planar Co(III) complexes bind two equivalents of a substrate whereas mechanistic studies
Accepted 9th December 2016
established the involvement of a Co3+–Co2+ based redox cycle in the catalysis. The Co(III) complex having
DOI: 10.1039/c6qi00389c electron-withdrawing –Cl substituents on the ligand was noted to be a much better catalyst than the one
rsc.li/frontiers-inorganic with electron-donating –CH3 substituents.

Introduction reduction of a nitro group; and (ii) to gain mechanistic


insights about such a critical reaction. Herein, we present two
Amines are important intermediates and key precursors in the square-planar Co3+ complexes supported with pyrrolecarboxa-
synthesis of various fine chemicals and pharmaceuticals and mide ligands and their role in the chemoselective reduction of
therefore developing efficient and versatile catalysts for the assorted nitro compounds encompassing aromatic, hetero-
selective reduction of a nitro group has always been significant cyclic, and aliphatic substrates using hydrazine hydrate. The
yet challenging.1,2 Traditionally, the reduction of a nitro group binding studies infer that the square-planar Co3+ complexes
has been achieved using various precious3 as well as transition can interact with two equivalents of a potential substrate
metal based catalysts.2 In most cases, precious metal based whereas mechanistic studies substantiate the involvement of a
catalysts are found to be efficient for the nitro reduction.3 Co3+–Co2+ based redox cycle in the catalysis.
Efforts have also been made to develop heterogeneous catalysts
including nano-materials4 and a few unconventional materials
such as activated carbon,5 functionalized carbon nanotubes,6
Results and discussion
graphene oxide,7 and assorted doped materials.8 Although
some of these alternative materials have been quite effective Synthesis and characterization
but (i) their cumbersome synthesis to obtain the actual cata- This work utilizes two ligands, N,N′-(4,5-dichloro-1,2-
lyst; (ii) requirement of harsh reaction conditions; and (iii) ill- phenylene)bis(1H-pyrrole-2-carboxamide) (H4L1) and N,N′-(4,5-
explained reaction mechanism significantly hamper their dimethyl-1,2-phenylene)bis(1H-pyrrole-2-carboxamide) (H4L2),
wider applicability. This reaction becomes even more challen- which differ by the placement of either –Cl or –Me substi-
ging when a nitro group is part of an aliphatic scaffold instead tuents on the phenylene ring.9 The reaction of the deproto-
of an arene ring;1 and for the substrates containing multiple nated form of the respective ligand with (Et4N)2[CoCl4] in DMF
functional groups.1,2 Therefore, it is eminently desirable followed by aerial oxidation affords Co(III) complexes 1 and 2
although challenging (i) to develop efficient and non-precious (Scheme 1). The disappearance of N–H stretches and a batho-
metal based less-expensive catalysts for the chemoselective chromically shifted O-amide band affirms the bonding
through the deprotonated ligands in both complexes (Fig. S1
and S2, ESI†).9,10 Complexes 1 and 2 are deep purple in color
Department of Chemistry, University of Delhi, Delhi – 110 007, India. and display prominent absorption features at ca. 560 and
E-mail: rgupta@chemistry.du.ac.in; http://people.du.ac.in/~rgupta/;
ca. 740 nm in CH3OH (Fig. S3, ESI†). The absorption spectra
Tel: +91-11-27666646
† Electronic supplementary information (ESI) available. CCDC 1445062. For ESI
recorded in CH3CN, CH3OH, and DMF displayed minor
and crystallographic data in CIF or other electronic format see DOI: 10.1039/ changes in the spectral features suggesting potential co-
c6qi00389c ordination of solvent molecule(s) in the solution state (Fig. S4

324 | Inorg. Chem. Front., 2017, 4, 324–335 This journal is © the Partner Organisations 2017
View Article Online

Inorganic Chemistry Frontiers Research Article

perfect square-planar and tetrahedral geometries, respect-


ively.11 The sum of angles around the Co3+ ion is 359.8°, pretty
close to 360°, thus again suggesting a predominantly square-
planar geometry. Notably, 1 represents a rare example of a
mono-nuclear square-planar Co(III) complex as only a handful
compounds are known in the literature.12 Co(III) complexes
typically prefer an octahedral geometry, however, in case of
strong σ-donation, the axial Lewis acidity can be reduced to
Published on 09 December 2016. Downloaded by Universitat de Barcelona on 8/27/2020 5:10:40 PM.

the point that a rare square-planar geometry becomes access-


ible as the case with complex 1.12b,d
Scheme 1 Synthetic routes for Co(III) complexes 1 and 2.
Solid-state variable-temperature (6–300 K) magnetic sus-
ceptibility studies exhibited that both complexes 1 and 2
follow the Curie–Weiss law over the temperature range of
60–300 K (Fig. S8, ESI†). At 300 K, μeff values of 3.2μB and 3.4μB
and S5, ESI†). It is therefore proposed that the solid-state four-
were observed for complexes 1 and 2, respectively. However,
coordinate Co(III) ion in complexes 1 and 2 is likely to interact
much reduced values of μeff both for 1 and 2 were noted at 6 K
with a suitable species and exceeds its coordination number,
suggesting the zero-field splitting.12b,13 These results are con-
at least in the solution state (vide infra). Both complexes are
sistent with the spin triplet (S = 1) ground state.12b,13 Both
sparingly soluble in H2O and therefore their absorption
complexes also displayed paramagnetically shifted 1H NMR
spectra could not be convincingly investigated; however,
spectra as expected for square-planar Co(III) complexes with
addition of excess H2O to a CH3OH solution of 1 and 2 did not
the S = 1 ground state.12b–d A representative spectrum of
result in an appreciable change (Fig. S6 and S7, ESI†).
complex 2, displayed in Fig. 2, clearly illustrates well-resolved
One of the representative compounds, complex 1, was struc-
protons resonating between 90 and −130 ppm. For example,
turally characterized. The crystal structure displays a nearly
pyrrole protons Hb, Hc, and Hd were noted as sharp reso-
square-planar geometry around the Co3+ ion wherein the
nances at −48, −13, and −127 ppm whereas arene-Ha was
ligand coordinates via two anionic Namide and two anionic
seen at 67 ppm. The –CH3 groups present on the arene ring
Npyrrole donors (Fig. 1). The average Co–Namide and Co–Npyrrole
were found to resonate at 89 ppm. A similar paramagnetically
bond distances were found to be 1.832 and 1.867 Å, respectively.9a,c
shifted 1H NMR spectrum was observed for complex 1
The five-membered metal-based chelate angles were smaller
(Fig. S9, ESI†).
than 90° therefore justifying a tight chelation around the
In cyclic voltammetric studies, both complexes display
cobalt center. The geometry around the Co3+ ion is a slightly
metal-based as well as ligand-based redox responses (Fig. 3).
distorted square-planar with a τ4 distortion parameter of
Complexes 1 and 2 exhibited reversible reductive responses at
0.15.11 For reference, τ4 has standard values of 0 and 1 for
−0.38 V (ΔE = 80 mV) and −0.48 V (ΔE = 85 mV), respectively
that has been assigned to the Co3+/2+ redox couple. In addition,
quasi-reversible oxidative responses at 0.30 V and 0.62 V for 1
and 0.14 V and 0.46 V for 2 have been tentatively suggested to
be ligand-centered. This assignment has been made by the
observation of oxidative responses in a similar range for our
earlier metal complexes supported with identical ligands.9a
Notably, the Co2+/+ redox couple was not observed within the

Fig. 1 Molecular structure of the anionic part of complex 1; thermal


ellipsoids are drawn at the 50% probability level whereas the Et4N+ ion
and hydrogen atoms have been omitted for clarity. Selected bond
distances (Å): Co–N1 = 1.881(7), Co–N2 = 1.837(7), Co–N3 = 1.828(7),
Co–N4 = 1.853(7). Selected bond angles (°): N1–Co–N2 = 85.1(3),
N1–Co–N3 = 169.1(3), N1–Co–N4 = 104.4(3), N2–Co–N3 = 84.7(3), Fig. 2 1H NMR spectrum of complex 2 recorded in CD3CN. * Represents
N2–Co–N4 = 170.0(3), N3–Co–N4 = 85.6(3). the residual solvent.

This journal is © the Partner Organisations 2017 Inorg. Chem. Front., 2017, 4, 324–335 | 325
View Article Online

Research Article Inorganic Chemistry Frontiers

supports both +3 as well as +2 oxidation states; (iii) contrast


electronic substituents (–Cl versus –CH3) could modulate the
electron density at the metal center as evidenced by the cyclic
voltammetric studies and such a fact may have a notable effect
on catalysis. Such structural and redox features advocated that
the present complexes are potential candidates for redox-
based catalysis.4c,16 We selected hydrazine-mediated nitro
reduction reactions to test the present complexes as cobalt has
Published on 09 December 2016. Downloaded by Universitat de Barcelona on 8/27/2020 5:10:40 PM.

recently emerged as an efficient catalytic metal in assorted


redox-based catalysis in general and nitro reduction in
particular.4c,16,17
An initial screening to identify the optimized reaction con-
ditions suggested that only 1 mol% of complex 1 was sufficient
Fig. 3 Cyclic voltammograms of complex 1 (green trace) and 2 (red
trace) recorded in MeOH at Pt working electrode; reference electrode: for the hydrazine-based reduction of nitrobenzene. However,
Ag+/Ag; auxiliary electrode: Pt wire; scan rate: 100 mV s−1; supporting for further optimization experiments, a challenging substrate
electrolyte: 0.1 M TBAP. having a –CN group, para-CNPhNO2 was chosen as a model
substrate and complex 1 as a representative catalyst (Table 1).
The presence of complex 1 as a catalyst was essential as
solvent window for both complexes. We believe the tetra- the absence or presence of other cobalt salts (CoCl2,
anionic ligands with high σ-donating abilities do not support (Et4N)2[CoCl4], or K3[Co(NO2)6]) as the catalysts did not result
the Co(I) state in the present complexes.9,12 It is important to in any reduction (entries 1 and 2, Table 1). Satisfyingly,
note that the electron-withdrawing –Cl groups in 1 have complex 1 quantitatively reduced para-CNPhNO2 in the pres-
respectively shifted Co3+/2+ and other oxidative responses by ence of hydrazine in MeOH at 60 °C (entry 3, Table 1).
ca. 100 and ca. 150 mV towards positive potentials. 9a,14,15 Hydrazine was found to be the most effective reducing agent
Thus, it could be inferred that the Co3+ ion in complex 1 could whereas H2, ammonium formate, NaBH4, formic acid, tri-
be easily reduced to Co2+ as a result of electron-withdrawing –Cl methylsilylsilane and triphenylsilane were mostly ineffective
substituents. although ca. 10% reduction was noted in the case of NaBH4
(entry 4). Entries 5 and 6 suggest the requirement of 60 °C as
Nitro reduction reactions the optimum temperature and 6 h duration for the quantita-
Both complexes 1 and 2 offered several noteworthy features tive reduction. Entry 7 supports that only 2 equivalents of
which suggested their potential importance in catalysis: (i) the NH2NH2 can quantitatively reduce para-CNPhNO2. Amongst
square-planar Co3+ ion suggests uncluttered access of a sub- different solvents screened (entries 8–10), MeOH, EtOH, H2O,
strate; (ii) electrochemical studies illustrate that the cobalt ion acetone, toluene, and THF resulted in nearly quantitative

Table 1 Optimization and control experiments using para-CNPhNO2 as a model substrate with hydrazine and using complex 1 as a representative
catalysta

S. no. Catalyst Reducing agent Conditions Yieldb (%)

1 — NH2NH2 60 °C 0
2 CoCl2 or (Et4N)2[CoCl4] or NH2NH2 60 °C 0
K3[Co(NO2)6]
3 1 NH2NH2 60 °C 100
4 1 H2,c HCOONH4, NaBH4, HCOOH, 60 °C 0, 0, 10, 0, 0, 0
(SiMe3)3SiH, Ph3SiH
5 1 NH2NH2 20, 30, 40, 50, 60 °C 25, 28, 30, 95, 100
6 1 NH2NH2 1, 2, 3, 4, 5, 6 h 20, 30, 50, 70, 90, 100
7 1 NH2NH2 0.5, 1, 2, 3 equiv. 40, 55, 99, 100
8 1 NH2NH2 MeOH, EtOH, i-PrOH, n-BuOH, t-BuOH 100, 100, 27, 60, 53
9 1 NH2NH2 H2O, acetone, MeCN 98, 100, 87
10 1 NH2NH2 Toluene, THF, CHCl3, dioxane 100, 96, 46, 42
a
Reaction conditions: catalyst: 1 mol%; time: 6 h; solvent: MeOH except for entries 8–10. b Yield was calculated using the gas chromatograph. c 1 atm.

326 | Inorg. Chem. Front., 2017, 4, 324–335 This journal is © the Partner Organisations 2017
View Article Online

Inorganic Chemistry Frontiers Research Article

reduction. Although water would have been the best solvent In the literature, there is a dearth of examples, except for a
but limited solubility of different nitro-substrates compelled few nanoparticle-based catalysts, for the reduction of aliphatic
us to use MeOH in the subsequent work. To rule out whether nitro compounds.4b–e,17 Recently, we have shown the hydrazine
in situ formed Co nanoparticles were actually acting as a cata- mediated reduction of a few aliphatic nitro substrates using tri-
lyst, a Hg drop test was performed.18 Complex 1 quantitatively nuclear complexes having catalytically active Co(II) centers.16
reduced para-CNPhNO2 to para-CNPhNH2 using hydrazine in To test the catalytic potential of the present Co(III) complexes,
the presence of a drop of mercury(0) thus eliminating the the reduction of a few aliphatic nitro compounds was
involvement of nanoparticles during the catalysis. attempted. Notably, 1 mol% catalyst loading was not effective
Published on 09 December 2016. Downloaded by Universitat de Barcelona on 8/27/2020 5:10:40 PM.

The optimised reaction conditions were recruited for the for a reasonable reduction of the aliphatic nitro substrates.
reduction of assorted nitro substrates (Table 2). Several However, on increasing the catalyst loading to 5 mol%,
aromatic nitro compounds having a variety of substituents complex 1 caused nearly quantitative reduction of nitrocyclo-
such as –Me, –F, –Cl, –CN, –NH2, and –COOEt (entries 2–9, pentane as well as tert-nitrobutane (96–98%, entries 18 and 19).
Table 2) were conveniently reduced to their corresponding As noted with earlier substrates, complex 2 was not quite
amines in a high yield. In particular, complex 1 resulted in effective even with 5 mol% catalyst loading. Notably, both
nearly quantitative reduction of most of the substrates complexes 1 and 2 were unsuccessful in reducing 3,5-dimethyl-
and such a fact clearly suggests its catalytic supremacy 4-nitrobenzoic acid (entry 21) whereas inconsequential
over complex 2. The catalysis scope was further extended reduction was noted for 2,6-dimethylnitrobenzene (entry 22).
to heterocyclic nitro compounds. For example, substrates, We believe that the presence of methyl groups on both sides of
such as 4-nitroimidazole, 5-nitroindole, and 5-nitro-2-amino- a nitro group in these substrates has restricted their access to
thiazole, were quantitatively reduced to their corresponding interact with the cobalt ion by creating enough steric crowding
amines (entries 10–12). The reduction of substrates contain- (vide infra).
ing two nitro groups often presents a challenge due to the It is important to note that the catalytic results are very
observation of incomplete reduction. Satisfyingly, complex 1 impressive for complex 1 as nearly quantitative reduction was
selectively and effortlessly reduced both nitro groups in 1,3- observed for most of the substrates. On the other hand,
dinitrobenzene, 1,4-dinitrobenzene, and 1,5-dinitronaphthal- complex 2 did not result in complete reduction in most cases.
ene (entries 13–15, Table 2). Interestingly, complex 2, on Such a fact strongly suggests a significant difference between
the other hand, always resulted in incomplete reduction the two complexes. Of course, the two complexes differ by
whereas two products, 1-amino-5-nitronaphthalene (27%) and having either electron-withdrawing –Cl groups on the o-phenyl-
1,5-diaminonaphthalene (55%) were obtained in the case of ene ring in complex 1 and electron-donating –CH3 substitu-
1,5-dinitronaphthalene. ents in complex 2. The electrochemical studies have ade-
To evaluate the selectivity of the two catalysts towards sub- quately justified the sizable redox difference between the two
strates offering two delicate functional groups, 3-nitrostyrene complexes. We suggest that the electron-withdrawing –Cl
(entry 16) and β-4-dinitrostyrene (entry 17) were employed.19 In groups in complex 1 reduce the electron density at the cobalt
the case of 3-nitrostyrene, 3-nitroethylbenzene was not center, therefore, de-stabilize the Co3+ state or favour its
observed with both complexes. Such a fact supports a good reduction to the Co2+ state. In other words, complex 1 is better
control over the chemoselectivity. Interestingly, complex 1 pro- suited to activate hydrazine to a large extent when compared
duced two products: 3-aminostyrene as the major product to complex 2. Such a stark difference in the reactivity between
(80%) and 3-aminoethylbenzene as the minor one (20%). two complexes is significant and illustrates the importance of
A somewhat similar situation was observed with complex 2. It judiciously selected substituents in catalyst design.
appears that the formation of 3-aminoethylbenzene is due to
the activation of a double bond by the reduction-generated Mechanistic insights
–NH2 group. This hypothesis is backed by the fact that both In order to gain insight into the mechanism of reduction reac-
complexes do not produce 3-nitroethylbenzene at all. tion, absorption spectral studies were performed. To first
Furthermore, both complexes do not reduce styrene to ethyl- evaluate whether the present Co3+ complexes, offering a four-
benzene even with excess hydrazine (entry 20). This hypothesis coordinate square-planar geometry, show any tendency to bind
is further strengthened by considering the reduction of a potential substrate, substrate binding studies were carried
β-4-dinitrostyrene having two different types of nitro groups: out. For such studies, both complexes 1 and 2 were titrated
one directly attached to the benzene ring while the second one with 1-methyl-4-nitrobenzene and aniline as the potential sub-
connected through an ethenyl fragment. The reduction of strate and product, respectively, and the process was moni-
β-4-dinitrostyrene afforded two products: 4-(2-aminovinyl) tored spectroscopically. Importantly, both complexes 1 and 2
aniline as the major and 4-(2-nitrovinyl)aniline as the minor unambiguously showed the binding of two equivalents of
one without affecting the double bond. Collectively, these either 1-methyl-4-nitrobenzene or aniline (Fig. 4 and S10–S12,
results justify that the observation of 3-aminoethylbenzene in ESI†). The binding studies were further analyzed by the linear
the reduction of β-4-dinitrostyrene was indeed due to the acti- regression fitting for 1 : 2 binding between the Co3+ complexes
vation of the double bond and therefore the present catalytic to that of 1-methyl-4-nitrobenzene and aniline. The binding
results exhibit high chemo-selectivity. constants (K, 10−3 M−2) for 1-methyl-4-nitrobenzene were

This journal is © the Partner Organisations 2017 Inorg. Chem. Front., 2017, 4, 324–335 | 327
View Article Online

Research Article Inorganic Chemistry Frontiers

Table 2 Catalytic reduction of assorted nitro compounds using hydrazine in the presence of complexes 1 and 2 as the catalystsa

Yieldb (%)

S. no. Substrate Product 1 2

1 >99 75
Published on 09 December 2016. Downloaded by Universitat de Barcelona on 8/27/2020 5:10:40 PM.

2 >99 54

3 >99 51

4 >99 61

5 99 79

6 >99 89

7 98 42

8 >99 88

9 >99 89

10 >99 87

11 >99 84

12 >99 92

13 >99 83

14 >99 100

328 | Inorg. Chem. Front., 2017, 4, 324–335 This journal is © the Partner Organisations 2017
View Article Online

Inorganic Chemistry Frontiers Research Article

Table 2 (Contd.)

Yieldb (%)

S. no. Substrate Product 1 2

15 0 27
Published on 09 December 2016. Downloaded by Universitat de Barcelona on 8/27/2020 5:10:40 PM.

>99 55

16 0 0

80 54

20 24

17 40 32

60 50

18 98 54

19 96 51

20 0 0

21 0 0

22 12 8

a
Conditions: catalyst: 1 mol%; solvent: MeOH; reducing agent: hydrazine (2 equiv.); temperature: 60 °C, time: 6 h. b Yield was calculated using
the gas chromatograph. c Catalyst: 5 mol%.

This journal is © the Partner Organisations 2017 Inorg. Chem. Front., 2017, 4, 324–335 | 329
View Article Online

Research Article Inorganic Chemistry Frontiers


Published on 09 December 2016. Downloaded by Universitat de Barcelona on 8/27/2020 5:10:40 PM.

Fig. 5 UV-Vis spectral titration of complex 1 with hydrazine in MeOH;


and change in absorption intensity as a function of equivalents of hydra-
Fig. 4 UV-Vis titration of complex 1 with aniline in MeOH. Top inset:
zine (inset).
Change in absorbance as a function of moles of aniline. Bottom inset:
Linear regression fitting curve for 1 : 2 binding between complex 1 and
aniline.
to pale-brown species, with feature-less absorption spectra,
after one-electron reduction. Fig. 6 presents a controlled poten-
found to be 10.4 and 8.9 for complexes 1 and 2, respectively.20 tial electrolysis experiment for complex 2 that clearly demon-
Similarly, aniline displayed K values (10−3 M−2) of 10.8 and 12.0 strates the reduction of the Co3+ state to Co2+ after the transfer
for complexes 1 and 2, respectively. The binding constants, of one electron (coulomb count: 0.92). In essence, both sub-
although small, justify that both Co3+ complexes are able to strate binding studies and hydrazine based reduction monitor-
bind two equivalents of a potential substrate or a product. ing assert that the present Co3+ complexes 1 and 2 not only
Interestingly, complex 1 illustrated much larger binding affinity bind two equivalents of a substrate but also activate hydrazine
for 1-methyl-4-nitrobenzene when compared to complex 2. Such via a Co3+–Co2+ redox process.
a fact suggests that complex 1 is comparatively more electron More importantly, pale-brown ‘reduced’ species could be
deficient than that of 2, most probably due to the presence of quantitatively oxidized to the original Co3+ complex by
electron-withdrawing substituents and therefore displays larger exposure to molecular oxygen.21 Such a fact is adequately sup-
binding affinity. This in turn supports the better catalytic per- ported by the observation of reversible Co3+/2+ redox couples at
formance of complex 1 in the reduction reactions. fairly negative potentials both for complexes 1 and 2.
We then focussed our attention on the reduction behaviour Importantly, such reversibility is better maintained for
of complexes 1 and 2 towards hydrazine. The reduction reac- complex 1 (Fig. 7) when compared to complex 2 (Fig. S16,
tion was spectroscopically monitored after the incremental ESI†). Such a noteworthy difference between two related
addition of hydrazine to the MeOH solution of Co(III) com-
plexes. The addition of hydrazine to complexes 1 and 2
resulted in an immediate color change from deep purple to
pale brown with disappearance of spectral features at ca. 560
and ca. 740 nm (Fig. 5 and S13, ESI†). Importantly, both spec-
tral titrations established that only 0.25 equivalent of hydra-
zine is required for the complete transformation (see insets of
Fig. 5 and S13, ESI†). It is important to note that one equi-
valent of hydrazine generates 4 electrons and 4 protons before
being converted to N2.17,19 Therefore, the observed stoichio-
metry of one Co(III) complex to that of 0.25 equivalent of hydra-
zine unambiguously confirms that the Co3+ state has been
reduced by one electron to the Co2+ state. Further evidence
came from the cyclic voltammetric experiments that showed
that the incremental addition of hydrazine to a MeOH solution
of 1 or 2 resulted in color change to pale brown accompanied
by the decrease in the current from the metal-based redox Fig. 6 Spectro-electrochemical traces during the coulometric
reduction of complex 2; 2 (black trace) illustrates the original absorption
responses (Fig. S14 and S15, ESI†). Furthermore, controlled spectrum whereas traces A–G (multiple colours) were recorded during
potential electrolysis experiments also resulted in an identical the course of electron transfer. Inset displays the number of electrons
observation wherein deep-purple complexes 1 and 2 changed supplied as a function of current.

330 | Inorg. Chem. Front., 2017, 4, 324–335 This journal is © the Partner Organisations 2017
View Article Online

Inorganic Chemistry Frontiers Research Article

protons were provided by two cobalt complexes after the reac-


tion with hydrazine in a step-wise manner (Scheme 2b). Based
on the binding studies, it is proposed that the two substrates
first interact with a Co(III) complex followed by the sequential
electron and proton transfer leading to the product
(Scheme 2c). The reaction involves formation of different inter-
mediates in each step of the catalytic cycle, following either
pathway A or B.17,22 In pathway A, the first step involves the for-
Published on 09 December 2016. Downloaded by Universitat de Barcelona on 8/27/2020 5:10:40 PM.

mation of nitrosobenzene which is further reduced to


N-phenylhydroxylamine that ultimately leads to aniline. In
pathway B, nitrosobenzene and N-phenylhydroxylamine
combine to form azoxybenzene that reduces to azobenzene,
then to 1,2-diphenylhydrazine which finally proceeds to
aniline. To identify the intermediate(s) during the reaction,
reduction of nitrobenzene was monitored by UV/Vis spectral
Fig. 7 Recyclability experiment of complex 1 with respect to hydrazine;
(Fig. S17, ESI†) as well as 1H NMR spectral studies (Fig. S18,
original absorption spectrum of complex 1 (black trace A) followed by
reduction with 0.25 equiv. NH2NH2 further followed by oxidation with ESI†) at different time intervals. During such a reaction invol-
O2 (traces B–J). Inset displays subsequent four regenerative cycles. ving complex 1 as a catalyst, N-phenylhydroxylamine was
detected as an intermediate in the absorption spectrum as
well as 1H NMR spectrum after 2 h. In addition, the generation
compounds suggests that while complex 1 maintains its struc- of N-phenylhydroxylamine was further confirmed by recording
tural integrity during the hydrazine-based reduction and the high-resolution mass spectrum displaying the M + H+ peak
oxygen-mediated oxidation; complex 2 is probably not that at 110.0607 (Fig. S19, ESI†). These experiments suggest that
stable. In fact, catalytic results are in perfect sync with these pathway A was operational during the reduction. To further
observations. Nevertheless, these experiments established that support pathway A, we performed individual reduction of
the present cobalt complexes function as the redox catalysts nitrosobenzene as well as N-phenylhydroxylamine; both such
for the hydrazine mediated reduction of nitro substrates. reactions quantitatively produced aniline. However, the
The hydrazine-mediated reduction typically proceeds in reduction of azobenzene as well as 1,2-diphenylhydrazine did
three steps requiring two-electrons and two-protons at each not lead to aniline, therefore, firmly ruling out pathway B.
step (Scheme 2a). We suggest that two-electrons and two- To further provide a convincing chemical support to the
proposed mechanism (Scheme 2), we attempted the reduction
of 2,2′-dinitrobiphenyl.17b As shown in Scheme 3, 2,2′-dinitro-
biphenyl provides an effective way to trap some of the inter-
mediates being formed during the course of the reaction. With
complex 1, this reaction resulted in the formation of three pro-
ducts: the expected 2,2′-diaminebiphenyl in 44% yield and
benzo[c]cinnoline (20%) as well as benzo[c]cinnoline-5-oxide
(36%). While the formation of 2,2′-diaminebiphenyl was as
anticipated, the other two products justify the intermediacy of

Scheme 2 Proposed mechanism for the hydrazine mediated reduction of Scheme 3 Hydrazine based reduction of 2,2’-dinitrobiphenyl and its
nitroaromatics. For simplicity, the case of nitrobenzene has been illustrated. various products observed during the catalysis.

This journal is © the Partner Organisations 2017 Inorg. Chem. Front., 2017, 4, 324–335 | 331
View Article Online

Research Article Inorganic Chemistry Frontiers

‘nitrosobenzene’ and ‘N-phenylhydroxylamine’ during the mixture was added to a solution of (Et4N)2[CoCl4] (0.127 g,
reduction reaction which have condensed to produce benzo[c] 0.275 mmol) in 5 mL DMF under an inert atmosphere. The
cinnoline as well as benzo[c]cinnoline-5-oxide (Scheme 3). It is resulting mixture was further stirred for 1 h under an O2
worth mentioning that N-phenylhydroxylamine was observed atmosphere at room temperature which resulted in a dark
as an intermediate during the reduction of nitrobenzene both purple solution. This solution was filtered through a pad of
in UV/Vis spectral (Fig. S17, ESI†) and 1H NMR spectral studies celite in a medium porosity frit. The solvent was removed
(cf. Fig. S18, ESI†). On the other hand, complex 2 produced under reduced pressure and the crude product was isolated
benzo[c]cinnoline and benzo[c]cinnoline-5-oxide in 40 and after washing repeatedly with diethyl ether. Vapor diffusion of
Published on 09 December 2016. Downloaded by Universitat de Barcelona on 8/27/2020 5:10:40 PM.

30% yield respectively, while 2,2′-diaminebiphenyl was pro- diethyl ether to a MeCN solution of the crude product at room
duced in only a 10% yield. The larger proportions of benzo[c] temperature afforded a crystalline product within 2–3 d. Yield:
cinnoline and benzo[c]cinnoline-5-oxide compared to 2,2′- 0.098 g (64%). Anal. Calcd for C24H28N5O2Cl2Co (548.35):
diaminebiphenyl further justify the inferior nature of complex C 52.57, H 5.15, N 12.77; Found C 52.18, H 5.35, N 12.78. FTIR
2 as a catalyst in the nitro reduction reaction. spectrum (Zn–Se ATR, cm−1): 2990, 2927, 1605, 1595, 1537,
1457, 1001, 747. Conductivity (DMF, ca. 1 mM solution, 298 K):
Recyclability experiments ΛM = 85 Ω−1 cm2 mol−1. UV/Vis spectrum [λmax, nm (ε, M−1 cm−1),
We then focused on understanding the potential recyclability MeOH]: 742 (260), 560 (700), 495 (sh, 520). UV/Vis spectrum
of complexes 1 and 2 in catalysis. For such a purpose, two [λmax, nm (ε, M−1 cm−1), CH3CN]: 810 (sh, 250), 689 (sh, 330),
experiments were performed. In the first experiment, a fixed 545 (780), 487 (sh, 620), 383 (1000). UV/Vis spectrum [λmax,
amount of complex 1 (1 mol%) was taken in a reaction flask nm (ε, M−1 cm−1), DMF]: 833 (sh, 250), 678 (sh, 370),
while fresh batches of para-CNPhNO2 (1 equiv.) and hydrazine 546 (860), 489 (sh, 690), 384 (1150). MS (ESI−): m/z calcd for
(2 equiv.) were added. Importantly, under these conditions, [C16H8Cl2N4O2Co]− 416.9356, found 416.9375.
complex 1 was successful in reducing several batches of para- (Et4N)[CoL2] (2). This complex was synthesized in an identi-
CNPhNO2 with only ca. 5% drop in performance in the fifth cal manner as that of 1 using following reagents: H4L2 (0.10 g,
cycle (Fig. S20, ESI†). In the second experiment, the solvent 0.310 mmol), NaH (0.029 g, 1.240 mmol), (Et4N)2[CoCl4]
was removed from the reaction mixture followed by the (0.143 g, 0.310 mmol). Yield: 0.094 g (60%). Anal. Calcd for
addition of ethyl acetate which allowed the isolation of the C26H36N5O3Co (including 1H2O as observed in microanalysis
catalyst. The recovered complex 1 was then used in the next data, 525.52): C 59.42, H 6.90, N 13.33; Found C 58.98, H 6.88,
catalytic run without any further purification and/or regener- N 13.20. FTIR spectrum (Zn–Se ATR, cm−1): 3378, 2987,
ation. The same strategy was repeated for five consecutive runs 2945, 1663, 1608, 1592, 1548, 1478, 1444, 1390, 1000, 747.
(Fig. S21, ESI†). In such an experiment, approximately 2–4% Conductivity (DMF, ca. 1 mM solution, 298 K): ΛM = 80
loss in activity per catalytic run was noted which suggests that Ω−1 cm2 mol−1. UV/Vis spectrum [λmax, nm (ε, M−1 cm−1),
a part of the complex may have decomposed during every cata- MeOH]: 744 (185), 561 (480), 499 (sh, 370). UV/Vis spectrum
lytic cycle in addition to loss in recovery. Furthermore, at the [λmax, nm (ε, M−1 cm−1), CH3CN]: 927 (sh, 210), 828 (sh, 255),
end of the 2nd cycle, recovered complexes 1 and 2 were charac- 689 (sh, 360), 554 (660), 474 (sh, 480), 392 (sh, 740). UV/Vis
terized by the FTIR spectra (Fig. S22 and S23, ESI†) as well as spectrum [λmax, nm (ε, M−1 cm−1), DMF]: 929 (sh, 250), 834
by the X-ray powder diffraction (XRPD) studies (Fig. S24 and (sh, 330), 680 (sh, 450), 559 (860), 480 (sh, 660), 393 (1010). MS
S25, ESI†). While complex 1 displayed an excellent match to (ESI−): m/z calcd for [C24H12Cl2N4O2Co]− 377.044; Found 377.039.
that of the as-synthesized sample, recovered complex 2 did
Procedure for reduction of nitro substrates
show certain differences particularly in the XRPD pattern.
Such a fact justifies the recyclability experiments (using In a Schlenk reaction flask, the nitro substrate (1.0 mmol) was
NH2NH2 and O2) suggesting the limited stability of complex 2 taken in MeOH (2 ml) and the catalyst (1 mol%) was added.
and therefore its inferior catalytic performance. The content was stirred for 5 min followed by the addition of
hydrazine monohydrate (2.0 mmol) under inert conditions.
The reaction mixture was stirred at 60 °C for 6 h while progress
Experimental of the reaction was monitored by the gas chromatograph (GC).
After completion, the reaction mixture was passed through a
Materials and methods plug of 100–200 mesh silica and concentrated to afford the
All reagents were obtained from commercial sources and used organic product(s). The product(s) was identified and/or
as received. Solvents were dried and/or purified using standard characterized by the GC, GC-MS, and 1H NMR spectra wher-
literature methods.23 (Et4N)2[CoCl4] salt was synthesized ever required.
according to a literature method.24 Ligands H4L1 and H4L2
were synthesized according to our earlier report.9a Characterization data for a few representative products
(Et4N)[CoL1] (1). Ligand H4L1 (0.10 g, 0.275 mmol) was dis- Aniline. 1H NMR (400 MHz, CDCl3): δ 7.26–7.19 (m, 6H),
solved in 5 mL DMF and treated with solid NaH (0.026 g, 6.83 (t, J = 7.9 Hz, 3H), 6.73 (d, J = 8.0 Hz, 6H), 3.62 (s, 7H).
1.101 mmol) under the dinitrogen atmosphere. The mixture 13
C NMR (100 MHz, CDCl3): δ 146.76 (C1), 129.54 (C2), 118.69
was stirred until H2 evolution ceased (ca. 10 min). This (C3), 115.37 (C4).

332 | Inorg. Chem. Front., 2017, 4, 324–335 This journal is © the Partner Organisations 2017
View Article Online

Inorganic Chemistry Frontiers Research Article

4-Chloroaniline. 1H NMR (400 MHz, CDCl3): δ 7.09 (d, J = CCD diffractometer having an Xcalibur sapphire diffraction
8.6 Hz, 1H), 6.59 (d, J = 8.5 Hz, 1H), 3.54 (s, 1H). 13C NMR measurement device at 293(2) K using graphite-mono-
(100 MHz, CDCl3): δ 144.92 (C1), 129.21 (C3), 123.21 (C4), chromated Mo-Kα radiation (λ = 0.71073 Å).27 The empirical
116.35 (C2). absorption correction was applied using the spherical harmo-
4-Iodoaniline. 1H NMR (400 MHz, CDCl3): δ 7.40 (d, J = nics implemented in the SCALE3 ABSPACK scaling algo-
8.8 Hz, 1H), 6.45 (d, J = 8.6 Hz, 1H). 13C NMR (100 MHz, rithm.27 The structure was solved by the direct methods using
CDCl3): δ 139.07 (C4), 129.42 (C2), 120.07 (C3), 114.80 (C1). SIR-92 28 and refined by full-matrix least-squares refinement
Ethyl-4-aminobenzoate. 1H NMR (400 MHz, CDCl3): δ 7.84 techniques on F2 using SHELXL97.29 All calculations were
done using the Wingx crystallographic package.30 The
Published on 09 December 2016. Downloaded by Universitat de Barcelona on 8/27/2020 5:10:40 PM.

(d, J = 8.3 Hz, 4H), 6.62 (d, J = 8.5 Hz, 4H), 4.30 (q, J = 7.1 Hz,
4H), 3.96 (s, 3H), 1.35 (t, J = 7.3 Hz, 6H). 13C NMR (100 MHz, crystallographic data collection and structure solution para-
CDCl3): δ 166.87 (s), 150.78 (s), 131.65 (s), 120.17 (s), 113.88 (s), meters are presented in Table S1 (ESI†).
60.45 (s), 14.52 (s).
3-Nitroaniline. 1H NMR (400 MHz, CDCl3): δ 7.57 (d, J =
8.2 Hz, 1H), 7.49 (s, 1H), 7.28 (d, J = 8.2 Hz, 1H), 6.94 (d, J = Conclusions
8.0 Hz, 1H), 3.98 (s, 2H). 13C NMR (1001 MHz, CDCl3):
δ 149.28 (C1), 147.57 (C5), 130.02 (C3), 120.78 (C2), 113.20 This work has discussed the synthesis and characterization of
(C4), 109.09 (C6). two square-planar Co(III) complexes supported with pyrrole-
4-Nitroaniline. 1H NMR (400 MHz, CDCl3): δ 7.91 (d, J = carboxamide ligands. Both complexes were successfully used
9.1 Hz, 1H), 6.69 (s, 1H), 6.56 (d, J = 9.1 Hz, 1H). 13C NMR in the catalytic reduction of assorted nitro compounds encom-
(100 MHz, CDCl3): δ 156.22 (C1), 136.13 (C4), 126.93 (C3), passing aromatic, heterocyclic, and even aliphatic substrates.
112.89 (C4). Binding studies asserted that four-coordinated Co(III) com-
1,3-Phenylenediamine. 1H NMR (400 MHz, CDCl3): δ 6.94 plexes bind two equivalents of nitro substrates. The absorption
(t, J = 7.9 Hz, 2H), 6.11 (d, J = 7.8 Hz, 4H), 6.02 (s, 2H), 3.45 spectral titrations established that the mechanism involved
(s, 9H). 13C NMR (100 MHz, CDCl3): δ 147.60 (C1), 130.31 (C3), redox shuttling between Co3+ and Co2+ states adequately sup-
106.09 (C2), 102.04 (C4). ported by the electrochemical studies. Out of the two com-
1,4-Phenylenediamine. 1H NMR (400 MHz, CDCl3): δ 6.56 plexes, the one with electron-withdrawing substituents
(s, 4H), 3.31 (s, 4H). 13C NMR (100 MHz, CDCl3): δ 138.66 (C1), resulted in nearly quantitative reduction for most of the sub-
116.83 (C2). strates. Such a fact has been related to better stabilization of
the Co2+ state and therefore better activation of hydrazine.
Physical measurements These results illustrate the significance of well-defined coordi-
The conductivity measurements were done in DMF using the nation complexes in delineating a challenging reaction and
digital conductivity bridge from the Popular Traders, India future work is directed to explore similar complexes in other
(model number: PT-825). The elemental analysis data were noteworthy organic reactions.
obtained from the Elementar Analysensystem GmbH Vario EL-III
instrument. The NMR measurements were done using a Jeol
(400 MHz) instrument. The infra-red spectra (either as KBr pellet Acknowledgements
or Zn–Se ATR) were recorded using either Perkin-Elmer FTIR RG acknowledges the financial support from the Council of
2000 or Spectrum-Two spectrometers. The absorption spectra Scientific & Industrial Research (CSIR), New Delhi and the
were recorded using the Perkin-Elmer Lambda-25 spectrophoto- University of Delhi. Authors thank the CIF-USIC of this
meter. The cyclic voltammetric experiments were performed university for the instrumental facilities including X-ray data
using a CH Instruments electrochemical analyzer (model collection and AIRF of JNU, New Delhi for the magnetic
1120A). The cell contained a glassy-carbon electrode, a Pt wire measurements. SK thanks CSIR for the RA fellowship.
auxiliary electrode, and Ag+/Ag as the reference electrode.25 The
solutions were ca. 1 mM in the complex and ca. 0.1 M in the sup-
porting electrolyte, TBAP.25 Gas chromatographic experiments Notes and references
were carried out with the Perkin Elmer Clarus 580 with an Elite-
5 column. The variable-temperature magnetic susceptibility 1 (a) N. Ono, The Nitro Group in Organic Synthesis,
measurements were performed by using the PPMS (Cryogenics Wiley-VCH, New York, 2001; (b) A. M. Tafesh and
Ltd, USA) at 0.5 Tesla applied magnetic field and having a closed J. Weiguny, Chem. Rev., 1996, 96, 2035; (c) H.-U. Blaser,
cycle based cryogen free system. The susceptibilities were cor- C. Malan, B. Pugin, F. Spindler, H. Steiner and M. Studer,
rected for diamagnetic corrections by using literature values.26 Adv. Synth. Catal., 2003, 345, 103; (d) H.-U. Blaser,
H. Steiner and M. Studer, ChemCatChem, 2009, 1, 210.
Crystallography 2 (a) Y. J. Jang, S. Kim, S. W. Jun, B. H. Kim, S. Hwang,
Single crystals suitable for the X-ray diffraction studies were I. K. Song, B. M. Kim and T. Hyeon, Chem. Commun., 2011,
grown by the vapor diffusion of diethyl ether to a MeCN solu- 47, 3601; (b) V. Pandarus, R. Ciriminna, F. Béland and
tion of complex 1. Intensity data were collected on an Oxford M. Pagliaro, Adv. Synth. Catal., 2011, 353, 1306; (c) T. K. Yu,

This journal is © the Partner Organisations 2017 Inorg. Chem. Front., 2017, 4, 324–335 | 333
View Article Online

Research Article Inorganic Chemistry Frontiers

J. Zeng, B. K. Lim and Y. N. Xia, Adv. Mater., 2010, 22, 5188; 9 (a) S. Kumar, M. Munjal, J. Singh and R. Gupta,
(d) S. Zhao, H. Liang and Y. Zhou, Catal. Commun., 2007, 8, Eur. J. Inorg. Chem., 2014, 4957; (b) J. Singh, G. Hundal and
1305; (e) K. Zhu, M. P. Shaver and S. P. Thomas, Chem. Sci., R. Gupta, Eur. J. Inorg. Chem., 2009, 3259; (c) S. Yadav,
2016, 7, 3031; (f ) K. V. R. Chary and C. S. Srikanth, Catal. S. Kumar and R. Gupta, Eur. J. Inorg. Chem., 2015, 5534.
Lett., 2009, 128, 164; (g) X. B. Lin, M. Wu, D. Y. Wu, S. Kuga 10 K. Nakamoto, in Infrared and Raman Spectra of Inorganic
and T. E. Y. Huang, Green Chem., 2011, 13, 28; and Coordination Compounds, John Wiley & Sons, New York,
(h) W. W. Lin, H. Y. Cheng, J. Ming, Y. C. Yu and F. Y. Zhao, 1986.
J. Catal., 2012, 291, 149; (i) P. F. Luo, K. L. Xu, R. Zhang, 11 L. Yang, D. R. Powell and R. P. Houser, Dalton Trans., 2007,
Published on 09 December 2016. Downloaded by Universitat de Barcelona on 8/27/2020 5:10:40 PM.

L. Huang, J. Wang, W. H. Xing and J. Huang, Catal. Sci. 955.


Technol., 2012, 2, 301; ( j) Y. K. Park, S. B. Choi, H. J. Nam, 12 Selected examples: (a) R. E. Patterson, S. W. Gordon-Wylie,
D.-Y. Jung, H. C. Ahn, K. Choi, H. Furukawa and J. Kim, C. G. Woomer, R. E. Norman, S. T. Weintraub,
Chem. Commun., 2010, 46, 3086; (k) G. S. Samuelsen, C. P. Horwitz and T. J. Collins, Inorg. Chem., 1998, 37, 4748;
V. L. Garik and G. B. L. Smith, J. Am. Chem. Soc., 1950, 72, (b) T. J. Collins, T. G. Richmond, B. D. Santarsiero and
3872; (l) R. Dey, N. Mukherjee, S. Ahammed and B. G. R. T. Treco, J. Am. Chem. Soc., 1986, 108, 2088;
B. C. Ranu, Chem. Commun., 2012, 48, 7982; (c) T. J. Collins, R. D. Powell, C. Slebodnick and
(m) G. Wienhöfer, I. Sorribes, A. Boddien, F. Westerhaus, E. S. Uffelman, J. Am. Chem. Soc., 1991, 113, 8419;
K. Junge, H. Junge, R. Llusar and M. Beller, J. Am. Chem. (d) J. C. Brewer, T. J. Collins, M. R. Smith and
Soc., 2011, 133, 12875; (n) S. Park, I. S. Lee and J. Park, Org. B. D. Santarsiero, J. Am. Chem. Soc., 1986, 110, 423.
Biomol. Chem., 2013, 11, 395; (o) W. Y. Wang, D. S. Wang, 13 (a) E. A. Boudreaux and L. N. Mulay, Theory and
X. W. Liu, Q. Peng and Y. D. Li, Chem. Commun., 2013, 49, Applications of Molecular Paramagnetism, Wiley, New York,
2903; ( p) M. Baron, E. Métay, M. Lemaire and F. Popowycz, 1976, pp. 198–211; (b) O. Kahn, Molecular Magnetism, Wiley
Green Chem., 2013, 15, 1006; (q) U. Sharma, N. Kumar, VCH, New York, 1993.
P. K. Verma, V. Kumar and B. Singh, Green Chem., 2012, 14, 14 (a) S. K. Sharma, S. Upreti and R. Gupta, Eur. J. Inorg.
2289; (r) S. K. Mohapatra, S. U. Sonavane, R. V. Jayaram and Chem., 2007, 3247; (b) S. K. Sharma, G. Hundal and
P. Selvam, Tetrahedron Lett., 2002, 43, 8527. R. Gupta, Eur. J. Inorg. Chem., 2010, 621; (c) S. K. Sharma
3 (a) A. Corma and P. Serna, Science, 2006, 313, 332; and R. Gupta, Inorg. Chim. Acta, 2011, 376, 95; (d) S. Kumar
(b) L. He, L.-C. Wang, H. Sun, J. Ni, Y. Cao, H.-Y. He and and R. Gupta, Indian J. Chem., Sect. A, 2011, 50, 1369;
K.-N. Fan, Angew. Chem., Int. Ed., 2009, 48, 9538; (c) M. Li, (e) S. Kumar and R. Gupta, Eur. J. Inorg. Chem., 2014, 5567;
L. Hu, X. Cao, H. Hong, J. Lu and H. Gu, Chem. – Eur. J., (f) S. Kumar, R. R. Jha, S. Yadav and R. Gupta,
2011, 17, 2763; (d) H. Wu, L. Zhuo, Q. He, X. Liao and New J. Chem., 2015, 39, 2042.
B. Shi, Appl. Catal., A, 2009, 366, 44; (e) A. J. Amali and 15 (a) M. Munjal and R. Gupta, Inorg. Chim. Acta, 2010, 363,
R. K. Rana, Green Chem., 2009, 11, 1781; (f ) J. Li, X. Shi, 2734; (b) M. Munjal and R. Gupta, Inorg. Chim. Acta, 2011,
Y. Bi, J. Wei and Z. Chen, ACS Catal., 2011, 1, 657. 372, 266; (c) M. Munjal, S. Kumar, S. K. Sharma and
4 (a) Nanoparticles and Catalysis, ed. D. Astrus, Wiley-VCH, R. Gupta, Inorg. Chim. Acta, 2011, 377, 144.
Weinheim, Germany, 2008; (b) E. Boymans, S. Boland, 16 S. Srivastava, M. S. Dagur, A. Ali and R. Gupta, Dalton
P. T. Witte, C. Muller and D. Vogt, ChemCatChem, 2013, 5, Trans., 2015, 44, 17453.
431; (c) F. A. Westerhaus, R. V. Jagadeesh, G. Wienhöfer, 17 (a) R. K. Rai, A. Mahata, S. Mukhopadhyay, S. Gupta,
M.-M. Pohl, J. Radnik, A.-E. Surkus, J. Rabeah, K. Junge, P.-Z. Li, K. T. Nguyen, Y. Zhao, B. Pathak and S. K. Singh,
H. Junge, M. Nielsen, A. Brückner and M. Beller, Nat. Inorg. Chem., 2014, 53, 2904; (b) D. Cantillo,
Chem., 2013, 5, 537; (d) D. Cantillo, M. Baghbanzadeh and M. M. Moghaddam and C. O. Kappe, J. Org. Chem., 2013,
C. O. Kappe, Angew. Chem., Int. Ed., 2012, 51, 10190; 78, 4530.
(e) X. Gu, Z. Sun, S. Wu, W. Qi, H. Wang, X. Xu and D. Su, 18 G. H. Whitesides, M. Hackett, R. L. Brainard,
Chem. Commun., 2013, 49, 10088. J.-P. P. M. Lavalleye, A. F. Sowinski, A. N. Izumi, S. S. Moore,
5 (a) D. S. Su, S. Perathoner and G. Centi, Chem. Rev., 2013, D. W. Brwon and E. M. Staudt, Organometallics, 1985, 4, 1819.
113, 5782; (b) Y. Lin and D. Su, ACS Nano, 2014, 8, 7823. 19 (a) U. Sharma, P. Kumar, N. Kumar, V. Kumar and B. Singh,
6 (a) J. Long, X. Xie, J. Xu, Q. Gu, L. Chen and X. Wang, ACS Adv. Synth. Catal., 2010, 352, 1834; (b) R. V. Jagadeesh,
Catal., 2012, 2, 6223; (b) Y. Cao, H. Yu, J. Tan, F. Peng, G. Wienhöfer, F. A. Westerhaus, A.-E. Surkus, M.-M. Pohl,
H. Wang, J. Li, W. Zheng and N. B. Wong, Carbon, 2013, 57, H. Junge, K. Junge and M. Beller, Chem. Commun., 2011,
433; (c) S. Wu, G. Weng, R. Schlögl and D. S. Su, Phys. 47, 10972.
Chem. Chem. Phys., 2015, 17, 1567. 20 Equilibrium data were determined by UV/Vis spectral titra-
7 Y. Gao, D. Ma, C. Wang, J. Guan and X. Bao, Chem. tion studies for complex 1/2 + 2S ⇌ [(1/2)(S)2] in CH3OH;
Commun., 2011, 47, 2432. where S = 1-methyl-4-nitrobenzene or aniline. For complex
8 (a) S. I. Fujita, H. Watanabe, A. Katagiri, H. Yoshida and 1, Keq: 104.17 mol−2 L2 for 1-methyl-4-nitrobenzene and
M. Arai, J. Mol. Catal. A: Chem., 2014, 393, 257; (b) Y. Lin, 92.83 mol−2 L2 for aniline. For complex 2, Keq:
S. Wu, W. Shi, B. Zhang, J. Wang, Y. A. Kim, M. Endo and 112.30 mol−2 L2 for 1-methyl-4-nitrobenzene and
D. S. Su, Chem. Commun., 2015, 51, 13086. 83.40 mol−2 L2 for aniline. See ref. 12b.

334 | Inorg. Chem. Front., 2017, 4, 324–335 This journal is © the Partner Organisations 2017
View Article Online

Inorganic Chemistry Frontiers Research Article

21 In case of electrochemical experiment, reversible oxidation 24 N. S. Gill and F. B. Taylor, Inorg. Synth., 1967, 9, 136.
to the initial Co(III) complex by O2 was hampered due to 25 (a) D. T. Sawyer and J. L. Roberts, in Experimental
the presence of excess supporting electrolyte, TBAP. Such a Electrochemistry for Chemists, Wiley, New York, 1974;
fact was independently confirmed by first reducing the (b) N. G. Connelly and W. E. Geiger, Chem. Rev., 1996, 96,
Co3+ complex using 0.25 equivalent of hydrazine followed 877.
by the addition of excess TBAP. Such a sequence did not 26 C. J. O’Connor, Prog. Inorg. Chem., 1982, 29, 203.
allow the regeneration of original Co3+ species by O2 27 CrysAlisPro, ver. 1.171.33.49b, Oxford Diffraction Ltd, 2009.
mediated oxidation. 28 A. Altomare, G. Cascarano, C. Giacovazzo and
Published on 09 December 2016. Downloaded by Universitat de Barcelona on 8/27/2020 5:10:40 PM.

22 (a) F. Z. Haber, Z. Elektrochem. Angew. Phys. Chem., 1898, A. Guagliardi, J. Appl. Crystallogr., 1993, 26, 343.
22, 506; (b) R. M. Dyson, M. Hazenkamp, K. Kaufmann, 29 G. M. Sheldrick, Acta Crystallogr., Sect. A: Fundam.
M. Maeder, M. Studer and A. J. Zilian, Chemometrics, 2000, Crystallogr., 2008, 64, 112.
14, 737. 30 L. J. Farrugia, WinGX version 1.64, An Integrated System of
23 D. D. Perrin, W. L. F. Armarego and D. R. Perrin, Windows Programs for the Solution, Refinement and Analysis
Purification of Laboratory Chemicals, Pergamon Press, of Single-Crystal X-ray Diffraction Data, Department of
Oxford, UK, 1980. Chemistry, University of Glasgow, 2003.

This journal is © the Partner Organisations 2017 Inorg. Chem. Front., 2017, 4, 324–335 | 335

You might also like