ArticlePDF Available

Copper(II) coordination polymers derived from triethanolamine and pyromellitic acid for bioinspired mild peroxidative oxidation of cyclohexane

Authors:

Abstract and Figures

The new inorganic 1D coordination polymer [Cu2(H3tea)2(mu4-pma)]n has been prepared, via self-assembly in aqueous medium, from copper(II) nitrate, triethanolamine (H3tea), pyromellitic acid (H4pma) and lithium hydroxide, and characterized by IR spectroscopy, elemental and single-crystal X-ray diffraction analyses. This compound and the related 2D polymer [Cu2(mu-H2tea)(2){mu3-Na2(H2O)4}(mu6-pma)]n.10nH2O are shown to mimic the alkane partial oxidation activity of the multicopper particulate methane monooxygenase, acting as catalysts precursors for the peroxidative oxidation of cyclohexane into cyclohexanol and cyclohexanone, by hydrogen peroxide (as green oxidant) and at room temperature in acidic MeCN/H2O medium. An overall yield (based on cyclohexane) of 29% has been achieved.
Content may be subject to copyright.
This article appeared in a journal published by Elsevier. The attached
copy is furnished to the author for internal non-commercial research
and education use, including for instruction at the authors institution
and sharing with colleagues.
Other uses, including reproduction and distribution, or selling or
licensing copies, or posting to personal, institutional or third party
websites are prohibited.
In most cases authors are permitted to post their version of the
article (e.g. in Word or Tex form) to their personal website or
institutional repository. Authors requiring further information
regarding Elsevier’s archiving and manuscript policies are
encouraged to visit:
http://www.elsevier.com/copyright
Author's personal copy
Copper(II) coordination polymers derived from triethanolamine
and pyromellitic acid for bioinspired mild peroxidative
oxidation of cyclohexane
Yauhen Y. Karabach
a
, Alexander M. Kirillov
a
, Matti Haukka
b
,
Maximilian N. Kopylovich
a
, Armando J.L. Pombeiro
a,*
a
Centro de Quı
´mica Estrutural, Complexo I, Instituto Superior Te
´cnico, TU Lisbon, Avenue Rovisco Pais, 1049–001 Lisbon, Portugal
b
University of Joensuu, Department of Chemistry, P.O. Box 111, FIN-80101 Joensuu, Finland
Received 19 September 2007; received in revised form 12 November 2007; accepted 19 November 2007
Available online 4 January 2008
Abstract
The new inorganic 1D coordination polymer [Cu
2
(H
3
tea)
2
(l
4
-pma)]
n
has been prepared, via self-assembly in aqueous medium, from
copper(II) nitrate, triethanolamine (H
3
tea), pyromellitic acid (H
4
pma) and lithium hydroxide, and characterized by IR spectroscopy, ele-
mental and single-crystal X-ray diffraction analyses. This compound and the related 2D polymer [Cu
2
(l-H
2
tea)
2
{l
3
-Na
2
(H
2
O)
4
}
(l
6
-pma)]
n
10nH
2
O are shown to mimic the alkane partial oxidation activity of the multicopper particulate methane monooxygenase,
acting as catalysts precursors for the peroxidative oxidation of cyclohexane into cyclohexanol and cyclohexanone, by hydrogen peroxide
(as green oxidant) and at room temperature in acidic MeCN/H
2
O medium. An overall yield (based on cyclohexane) of 29% has been
achieved.
Ó2007 Elsevier Inc. All rights reserved.
Keywords: Copper coordination polymers; Particulate methane monooxygenase; C–H activation; Hydrogen peroxide; Homogeneous catalysis
The search for new mild and efficient routes for
functionalization of alkanes to industrially valuable prod-
ucts constitutes a subject of high relevance in various areas
including environmental catalysis and green chemistry [1–
9]. A promising approach consists on the design and devel-
opment of new bioinspired catalytic materials [10–13].In
this respect, particulate methane monooxygenase (pMMO)
appears to be a unique copper enzyme that can catalyze the
hydroxylation of alkanes. It bears an active site composed
of a multinuclear Cu cluster possessing a NO-environment
[14–16]. However, although numerous examples of bioin-
spired copper compounds have been reported
[10–12,17,18], the synthetic and catalytic (with respect to
alkanes) studies on multinuclear copper complexes related
to pMMO remain rather scant [19–23].
Aiming at mimicking pMMO, we have recently reported
a series of multicopper compounds with NO-ligands and
applied them in the catalytic peroxidative oxidation of
alkanes [21–23]. In particular, the new aqua-soluble cop-
per–sodium 2D coordination polymer [Cu
2
(l-H
2
tea)
2
{l
3
-Na
2
(H
2
O)
4
}(l
6
-pma)]
n
10nH
2
O(1)(Scheme 1) has
been prepared [24] from triethanolamine (H
3
tea) and
pyromellitic acid (H
4
pma), and preliminary screened as a
potential bioinspired catalyst. In pursuit of that study,
the main objectives of the current work consist in (i) the
synthesis of another structurally distinct, although related,
copper coordination polymer, and (ii) a detailed evaluation
of the catalytic activity of both species towards the mild
peroxidative oxidation of cyclohexane to cyclohexanol
and cyclohexanone.
Hence, by following the previously reported [24] proce-
dure for 1, but adding lithium hydroxide instead of sodium
hydroxide as a pH regulator to an aqueous mixture
0162-0134/$ - see front matter Ó2007 Elsevier Inc. All rights reserved.
doi:10.1016/j.jinorgbio.2007.11.007
*
Corresponding author. Fax: +351 21 846 4455.
E-mail address: pombeiro@ist.utl.pt (A.J.L. Pombeiro).
www.elsevier.com/locate/jinorgbio
Available online at www.sciencedirect.com
Journal of Inorganic Biochemistry 102 (2008) 1190–1194
JOURNAL OF
Inorganic
Biochemistry
Author's personal copy
composed of copper(II) nitrate, triethanolamine and
pyromellitic acid, we have obtained, by self-assembly, the
polymeric compound [Cu
2
(H
3
tea)
2
(l
4
-pma)]
n
(2)(Scheme
1).
1
It has been isolated in ca. 50% yield as a blue air-stable
crystalline solid and characterized by IR spectroscopy,
elemental and single-crystal X-ray diffraction analyses.
2
A modification of the type of alkali metal hydroxide
(i.e. LiOH in 2vs. NaOH in 1) has resulted in a distinct
compound without incorporation of the alkali ions. Other
distinguishing features of 2consist in its insolubility in
common solvents, and in the presence of mononuclear
[Cu(H
3
tea)]
2+
subunits, in contrast to the good solubility
of 1in water and the dinuclear character of its [Cu
2
(l-
H
2
tea)
2
]
2+
core [24].
OH
Na
O
H2
H2
O
Na
OH2
OH2
O
Cu
N
HO
O
O
O
O
OO
OO
OH
O
Cu
NOH
1
1
2
23
3
4
4
(H2O)10
HO OH
Cu
N
HO
O
O
O
O
OO
OO
2
HO OH
Cu
N
HO
1
1
2
[Cu2(
µ
-H2tea)2{
µ
3-Na2(H2O)4
}(
µ
6-pma)]n·10nH2O (1) [Cu2(H3tea)2(
µ
4-pma)]n (2)
Scheme 1. Schematic representations (repeating units) of compounds 1and 2. Numbers indicate the corresponding extensions of polynuclear chains.
Fig. 1. Structural fragment of 2with the partial atom labelling scheme
and intramolecular hydrogen bonds (dashed lines). Various symmetry
transformations used to generate the equivalent non-hydrogen atoms.
Hydrogen atoms (apart from those involved in represented H-bonds) are
omitted for clarity. Displacement ellipsoids are drawn at the 30%
probability level. Cu green, N blue, O red, C grey. Selected bond lengths
(A
˚) and angles (°): Cu1–O1 1.945(3), Cu1–O3 2.365(3), Cu1–O4 1.992(3),
Cu1–O5 1.943(3), Cu1–O7 2.688(4), Cu1–N1 2.025(4), O1–Cu1–O5
89.20(13), O7–Cu1–O1 91.22(14), O4–Cu1–O5 91.78(13). Hydrogen bonds
DHA [d(DA) A
˚;\(DHA)°]: O4–H4O6 [2.587(4); 147.4], O7–
H7O2 [2.625(5); 156.2], O3–H3O7
i
[2.753(4); 171.8], symmetry code
(i): x+ 1/2, –y+ 1/2, z+ 1/2. (For interpretation of the references to color
in this figure legend, the reader is referred to the web version of this
article.)
1
Synthesis of [Cu
2
(H
3
tea)
2
(l
4
-pma)]
n
(2): To an aqueous solution
(10.0 mL) containing Cu(NO
3
)
2
2.5H
2
O (1.00 mmol) in HNO
3
(1.00 mmol) [the acid was added to avoid spontaneous hydrolysis of the
metal salt] were added dropwise triethanolamine (1.00 mmol, 130 lL), an
aqueous solution (3.0 mL) of LiOH (72 mg, 3.00 mmol) and pyromellitic
acid (0.50 mmol, 127 mg) in this order and with continuous stirring in air
at room temperature. The resulting reaction mixture was stirred overnight
and then filtered. The filtrate was left to evaporate in a beaker at ambient
(ca. 20–25 °C) temperature. Blue X-ray quality crystals in ca. 50% yield
(based on copper nitrate) were formed in one week, then collected and
dried in air. Anal. data for C
22
H
32
Cu
2
N
2
O
14
(675.6): Calc. C, 39.11; H,
4.77; N, 4.15%. Found: C, 38.74; H, 4.90; N 4.11%. IR (KBr pellet, band
type: vs. very strong; s, strong; m, medium; w, weak; br., broad; sh.,
shoulder): 3354 and 3264 s br. m(OH), 2992, 2899 and 2787 w m(CH), 1580
vs br. and 1494 m m
as
(COO), 1419 s (1385 sh.) and 1328 s (1286 sh.)
m
s
(COO), 1134 s, 1076 s, 1041 m, 1011 m, 900 s, 849 m, 802 m, 763 m, 680
w, 573 s, 541 m and 478 m (other bands) cm
1
. All chemicals were
obtained from commercial sources and used as received. All character-
ization procedures were performed on the instruments and according to
the techniques previously described [21,22]. Compound 1has been
obtained according to a published procedure [24].
2
X-ray crystal structure determination: An X-ray quality single crystal of
2was mounted in an inert oil within the cold dinitrogen stream of the
diffractometer. The X-ray diffraction data were collected with a Nonius
Kappa CCD diffractometer. The Denzo-Scalepack [32] program package
was used for cell refinements and data reduction. The structure was solved
by direct method using the SHELXS-97 program [33]. A multiscan
absorption correction based on equivalent reflections (XPREP in SHEL-
XTL) [34] was applied to all data. The structure was refined with
SHELXL-97 [35]. OH hydrogen atoms were located from the difference
Fourier map but not refined. Other H atoms were placed in idealized
positions and constrained to ride on their parent atoms. Crystal data for 2:
C
11
H
16
CuNO
7
,M= 337.79, monoclinic, space group P21/n,
a= 7.1237(4) A
˚,b= 14.5282(15) A
˚,c= 13.2247(14) A
˚,a=c= 90,
b= 104.139(6)°.V= 1327.2(2) A
˚
3
,Z=4, D
c
=1.690g/cm
3
,
l= 1.677 mm
1
,MoKaradiation, k= 0.71073 A
˚, 9915 reflections col-
lected, 2392 unique (R
int
= 0.0500), GOF = 1.077, R1 = 0.0453,
wR
2
= 0.1158, R1 = 0.0647 (all data), wR
2
= 0.1262 (all data).
Y.Y. Karabach et al. / Journal of Inorganic Biochemistry 102 (2008) 1190–1194 1191
Author's personal copy
The crystal structure of 2(Fig. 1) is composed of
repeated symmetry equivalent [Cu(H
3
tea)]
2+
units inter-
connected by centrosymmetric l
4
-pyromellitate(4–) anions,
which act as spacers thus forming ladder-like 1D polymeric
chains (Fig. 2, see also Supplementary data). The fully pro-
tonated H
3
tea moiety acts as a tetradentate N,O,O,O-che-
lating ligand filling a distorted tetragonal-bipyramidal
geometry around the Cu1 atom that is completed by two
pyromellitate oxygen atoms (O1 and O5). The binding of
H
3
tea involves three nonequivalent Cu–O bonds with dis-
tances of 1.945(3) A
˚[Cu1–O1], 2.365(3) A
˚[Cu1–O3] and
2.688(4) A
˚[Cu1–O7], the latter being significantly elon-
gated relatively to the average bond of that type but com-
parable to related bond lengths in other Cu compounds
derived from triethanolamine [21,22,24]. The linkage of
[Cu(H
3
tea)]
2+
and pma moieties is further stabilized by
strong intramolecular O4–H4O6 [2.587(4) A
˚] and O7–
H7O2 [2.625(5) A
˚] H-bonds between two triethanola-
mine OH groups and carboxylate C@O oxygen atoms,
leading to the formation of six-membered nonplanar
Cu–O–HO–C–O rings with the O4–Cu1–O5 and O7–
Cu1–O1 bite angles of 91.78(13)°and 91.22(14)°, respec-
tively. The separations between collateral copper centres
(all lying in one plane) within the polymeric chain depicted
in Fig. 2 are 7.124(1) A
˚and 8.446(1) A
˚, the former being
the aunit-cell dimension. That 1D chain contains cavities
formed by repeated sixteen-membered Cu–pma–Cu–pma
rings (i.e. A, B, C in Fig. 2) trough connective O1–Cu1–
O5 [89.20(13)°] angles (Fig. 1). In the crystal cell of 2,
the neighbouring metal–organic chains are held together
via multiple intermolecular H-bonds [O3–H3O7
i
2.753(4) A
˚], thus resulting in the formation of a 3D hydro-
gen bonded supramolecular assembly (Supplementary
data).
Both 1and 2are catalysts precursors for the peroxida-
tive oxidation of cyclohexane, by aqueous H
2
O
2
, to cyclo-
hexanol and cyclohexanone (Scheme 2).
3
The reaction
proceeds in MeCN/H
2
O, at room temperature and atmo-
spheric pressure, and in the presence of an acid additive
(HNO
3
). Cyclohexane has been chosen as a recognized
substrate model [25], also in view of the significance of
the products [25,26]. The effects on the catalytic activity
of various factors (HNO
3
,H
2
O
2
, catalyst and MeCN
amounts) have been tested for 1(Fig. 3,Table 1) aiming
at the optimization of the conditions. The HNO
3
quantity
has a significant effect (Fig. 3a), and the acid-to-catalyst
precursor molar ratio of 10:1 is sufficient for the observed
high activity, although more acid leads to slightly higher
overall yields. The increase of oxidant amount also pro-
vides higher yields (Fig. 3b), growing from 1.7 to 29% on
changing the oxidant-to-catalyst molar ratio from 125 to
1000. A similar behaviour is observed for 2, although with
a slight yield drop for the high oxidant content conceivably
due to overoxidation (Table 1, entries 3,4). The optimal
amounts of 1(Fig. 3c) and acetonitrile (Fig. 3d) are 10.0
lmol and 3.0–3.5 mL, respectively.
An interesting feature of the active species generated
from 1consists in the almost full preservation of activity
even after three catalyst recycles,
3
i.e. 26% vs. 29% (original
yield) (Table 1, entries 5–7). The activity of 1and 2(Table
1) is only slightly lower than that (39% yield) reported for
the most active copper catalysts in the mild cyclohexane
oxidation [21,22], and is much higher than those exhibited
by other copper systems [22], including a related coordina-
tion polymer derived from H
3
tea and terephthalic acid
[21,22]. With respect to pMMO, the activities of the com-
Fig. 2. Partial representation of the ladder-like polymeric chain of 2
(projection down the caxis). Hydrogen atoms are omitted for clarity. Cu
green, N blue, O red, C grey. Letters A, B, C correspond to the equivalent
sixteen-membered Cu–pma–Cu–pma rings. (For interpretation of the
references to color in this figure legend, the reader is referred to the web
version of this article.)
Cu catalyst precursor
aq. H2O2
MeCN/H2O, r.t.
OH O
+
Scheme 2.
3
Catalytic activity studies: The reaction mixtures were prepared as
follows: to 1.25–30.0 lmol (typically 10.0 lmol) of catalyst precursor
(compound 1or 2) contained in the reaction flask were added 0.0–10.0 mL
(typically 3.0–3.5 mL) MeCN, 0.0–0.50 mmol (typically 0.1 mmol) HNO
3
,
0.625 mmol of C
6
H
12
and 1.25–10.0 mmol H
2
O
2
(30% in H
2
O), in this
order. The reaction mixture was stirred for 6 h at room temperature (ca.
25 °C) and air atmospheric pressure, then 90 lL of cycloheptanone (as
internal standard), typically 10.0 mL diethyl ether (to extract the substrate
and the products from the reaction mixture) and 0.5 g PPh
3
(to reduce the
cyclohexyl hydroperoxide, according to a method developed by Shul’pin
[28]) were added. The resulting mixture was stirred for 15 min and then a
sample taken from the organic phase was analyzed by GC using a
FISONS Instruments GC 8000 series gas chromatograph with a DB WAX
fused silica capillary column (P/N 123-7032) and the Jasco-Borwin v.1.50
software. The GC analyses of the aqueous phase showed the presence of
only traces (less than 0.05%) of oxidation products. Catalyst recycling
experiments were performed as follows. On completion of each batch, the
products were analyzed as usually (except that PPh
3
was added directly to
the GC sample and not to all the reaction mixture) and the catalytic
species was recovered by full evaporation of the reaction mixture under
vacuum. The subsequent batches were initiated upon addition of new
standard portions of all the other reagents and solvent [22]. In the
experiments with radical traps, the appropriate compounds e.g. 2,2,6,6-
tetramethylpiperidine-1-oxyl (TEMPO), CBrCl
3
or Ph
2
NH (0.625 mmol)
were also added to the reaction mixture. Blank experiments were
performed with different amounts of H
2
O
2
and other reagents, and
confirmed that no cyclohexane oxidation products (or only traces, below
0.3%) were obtained in the absence of the metal catalyst.
1192 Y.Y. Karabach et al. / Journal of Inorganic Biochemistry 102 (2008) 1190–1194
Author's personal copy
pounds studied herein, based on mass and averaged over
the reaction time, i.e. 52 (for 1) and 72 (for 2) nmol of
products per min and mg of catalyst precursor (recalcu-
lated values from entries 1 and 2, Table 1), are higher than
that of pMMO although in different conditions, i.e. 17
nmol of EtOH per min and mg of protein, for the enzy-
matic hydroxylation of ethane, the most favourable sub-
strate for this enzyme [27].
Although the detailed mechanistic pathway is still to be
established, as for the other multinuclear copper catalysts
[21–23], it is believed to proceed through both C- and
O-centred radicals on account of the strong inhibition of
cyclohexane oxidation by the presence of traps for such
radicals (Supplementary data).
3
Cyclohexyl hydroperoxide
(CyOOH) is also detected (usually in ca. 3–5% yields, at the
end of the experiments)
3
as an intermediate [28]. The con-
version of CyOOH into the products can be metal-assisted
[22], conceivably involving the homolytic decomposition to
the alkyl peroxy CyOO
(upon O–H bond cleavage) and the
alkoxy CyO
(upon O–O bond rupture). Dismutation of
CyOO
would yield both cyclohexanol (CyOH) and cyclo-
hexanone with O
2
, while CyOH could also be derived upon
H-abstraction from cyclohexane (CyH) by CyO
[2,28–31].
Despite of being distinct in structural and solubility
properties, 1and 2exhibit similar levels of activity (maxi-
mal overall yields of ca. 29%) what suggests a similarity
of the active catalytic species. Further studies towards wid-
ening the family of pMMO inspired multicopper complexes
and extending their catalytic application from the partial
oxidation of cyclohexane to other alkanes are in progress.
Fig. 3. Effects of acid-to-catalyst (a) and oxidant-to-catalyst (b) molar ratios, catalyst (c) and acetonitrile (d) amounts on the total yields of cyclohexanol
and cyclohexanone (% relatively to C
6
H
12
) in the peroxidative oxidation of cyclohexane with catalyst precursor 1. In (c) TONs (dashed line) correspond to
moles of both products per mol of precursor 1. Reaction conditions: 6 h reaction time, room temperature, C
6
H
12
= 0.625 mmol, compound 1= 10.0 (a, b,
d) or 0.0–30.0 lmol (c), n(HNO
3
)/n(catalyst) = 10:1 (b–d) or HNO
3
= 0.0–0.50 mmol (a), H
2
O
2
= 10.0 (a, c, d) or 1.25–10.0 mmol (b), MeCN = 3.0 (b, c),
5.0 (a) or 0.0–10.0 mL (d).
Table 1
Peroxidative oxidation of cyclohexane to cyclohexanol and cyclohexanone
in the presence of compounds 1and 2
a
Entry Catalyst
precursor
n(H
2
O
2
)/
n(catalyst
precursor)
Yield of products, %
b
Cyclo-
hexanol
Cyclo-
hexanone
Total
c,d
111000 12.9 16.1 29.0
22500 17.4 10.5 27.9
3
e
2500 16.7 9.6 26.3
4
e
21000 14.4 8.8 23.2
511000 13.4 15.7 29.1
61(after run 5)
f
1000 11.4 15.3 26.7
71(after run 6)
f
1000 11.2 15.0 26.2
a
Selected data; reaction conditions (unless stated otherwise): C
6
H
12
(0.625 mmol), catalyst precursor (10 lmol), aqueous 30% H
2
O
2
(10.0 mmol), MeCN (3.5 mL for entries 1–4, or 3.0 mL for entries 5–7),
HNO
3
(0.10 mmol), 6 h reaction time, room temperature.
b
Moles of product/100 moles of C
6
H
12
.
c
Cyclohexanol + cyclohexanone.
d
Overall TON values (moles of products/moles of catalyst) are equal to
the total % yields (entries 3, 4), or can be estimated as [total %
yield] 0.625 (entries 1, 2, 5–7).
e
C
6
H
12
(1.00 mmol).
f
The catalytic species used in the previous reaction batch (run) was
recovered,
3
and the subsequent batch was initiated upon addition of new
standard portions of all the other reagents and solvent.
Y.Y. Karabach et al. / Journal of Inorganic Biochemistry 102 (2008) 1190–1194 1193
Author's personal copy
Acknowledgments
This work has been partially supported by the Founda-
tion for Science and Technology (FCT), Portugal, and its
POCI 2010 programme (FEDER funded), and a HRTM
Marie Curie Research Training Network (AQUACHEM
Project, CMTN-CT-2003-503864).
Appendix A. Supplementary data
Crystallographic data for the structure reported in this
paper have been deposited with the Cambridge Crystallo-
graphic Data Centre as supplementary publication No.
CCDC-661341 (2). Copies of the data can be obtained free
of charge on application to the CCDC, 12 Union Road,
Cambridge CB2 1EZ, UK (Fax: + 44 1223 336033; e-mail:
deposit@ccdc.cam.ac.uk or www: http://www.ccdc.cam.
ac.uk). Supplementary data associated with this article
can be found, in the online version, at doi:10.1016/
j.jinorgbio.2007.11.007.
References
[1] E.G. Derouane, F. Parmon, F. Lemos, F. Ramo
ˆa Ribeiro (Eds.),
Sustainable Strategies for the Upgrading of Natural Gas: Funda-
mentals, Challenges, and Opportunities, NATO Science Series, vol.
191, Kluwer Academic Publ., Dordrecht, The Netherlands, 2005.
[2] A.E. Shilov, G.B. Shul’pin, Activation and Catalytic Reactions of
Saturated Hydrocarbons in the Presence of Metal Complexes, Kluwer
Academic Publ., Dordrecht, The Netherlands, 2000.
[3] F.J.J.G. Janssen, R.A. van Santen (Eds.), Environmental Catalysis,
Imperial College Press, 1999.
[4] R.H. Crabtree, J. Organomet. Chem. 289 (2004) 4083–4091.
[5] J.A. Labinger, J.E. Bercaw, Nature 417 (2002) 507–514.
[6] A.A. Fokin, P.R. Schreiner, Adv. Synth. Catal. 345 (2003) 1035–1052.
[7] A.A. Fokin, P.R. Schreiner, Chem. Rev. 102 (2002) 1551–1593.
[8] C. Jia, T. Kitamura, Y. Fujiwara, Acc. Chem. Res. 34 (2001) 633–639.
[9] R.H. Crabtree, J. Chem. Soc., Dalton Trans (2001) 2437–2450.
[10] P. Gamez, P.G. Aubel, W.L. Driessen, J. Reedijk, Chem. Soc. Rev. 30
(2001) 376–385.
[11] L.M. Mirica, X. Ottenwaelder, T.D.P. Stack, Chem. Rev. 104 (2004)
1013–1045.
[12] E.A. Lewis, W.B. Tolman, Chem. Rev. 104 (2004) 1047–1076.
[13] D.S. Nesterov, V.N. Kokozay, V.V. Dyakonenko, O.V. Shishkin, J.
Jezierska, A. Ozarowski, A.M. Kirillov, M.N. Kopylovich, A.J.L.
Pombeiro, Chem. Commun. (2006) 4605–4607, and references
therein.
[14] R.L. Lieberman, A.C. Rosenzweig, Nature 434 (2005) 177–182.
[15] R.L. Lieberman, K.C. Kondapalli, D.B. Shrestha, A.S. Hakemian,
S.M. Smith, J. Telser, J. Kuzelka, R. Gupta, A.S. Borovik, S.J.
Lippard, B.M. Hoffman, A.C. Rosenzweig, T.L. Stemmler, Inorg.
Chem. 45 (2006) 8372–8381.
[16] S.I. Chan, V.C.-C. Wang, J.C.-H. Lai, S.S.-F. Yu, P.P.-Y. Chen,
K.H.-C. Chen, C.-Li Chen, M.K. Chan, Angew. Chem., Int. Ed 46
(2007) 1992–1994.
[17] S. Itoh, second ed., in: J.A. McCleverty, T.J. Meyer, L. Que, W.B.
Tolman (Eds.), Comprehensive Coordination Chemistry, vol. 8,
Elsevier, Dordrecht, 2003, pp. 369–393 (Chapter 8.15).
[18] D.H. Lee, second ed., in: J.A. McCleverty, T.J. Meyer, L. Que, W.B.
Tolman (Eds.), Comprehensive Coordination Chemistry, vol. 8,
Elsevier, Dordrecht, 2003, pp. 437–457 (Chapter 8.17).
[19] C. Shimokawa, J. Teraoka, Y. Tachi, S. Itoh, J. Inorg. Biochem. 100
(2006) 1118–1127.
[20] M.H. Groothaert, P.J. Smeets, B.F. Sels, P.A. Jacobs, R.A. Schoo-
nheydt, J. Am. Chem. Soc. 127 (2005) 1394–1395.
[21] A.M. Kirillov, M.N. Kopylovich, M.V. Kirillova, M. Haukka,
M.F.C.G. da Silva, A.J.L. Pombeiro, Angew. Chem., Int. Ed. 44
(2005) 4345–4349.
[22] A.M. Kirillov, M.N. Kopylovich, M.V. Kirillova, E.Y. Karabach, M.
Haukka, M.F.C.G. da Silva, A.J.L. Pombeiro, Adv. Synth. Catal. 348
(2006) 159–174, and references therein.
[23] C. Di Nicola, Y.Y. Karabach, A.M. Kirillov, M. Monari, L.
Pandolfo, C. Pettinari, A.J.L. Pombeiro, Inorg. Chem. 46 (2007)
221–230.
[24] Y.Y. Karabach, A.M. Kirillov, M.F.C.G. da Silva, M.N. Kopylo-
vich, A.J.L. Pombeiro, Cryst. Growth Des. 6 (2006) 2200–2203.
[25] U. Schuchardt, D. Cardoso, R. Sercheli, R. Pereira, R.S. da Cruz,
M.C. Guerreiro, D. Mandelli, E.V. Spinace, E.L. Pires, Appl. Catal.
A 211 (2001) 1–17.
[26] Ullmann’s Encyclopedia of Industrial Chemistry, Wiley-VCH, Wein-
heim, 6th ed., 2002.
[27] S.J. Elliott, M. Zhu, L. Tso, H.-H.T. Nguyen, J.H.-K. Yip, S.I. Chan,
J. Am. Chem. Soc. 119 (1997) 9949–9955.
[28] G.B. Shul’pin, J. Mol. Catal. A: Chem. 189 (2002) 39–66.
[29] M. Hartmann, S. Ernst, Angew. Chem., Int. Ed. 39 (2000) 888–890.
[30] M.V. Kirillova, A.M. Kirillov, P.M. Reis, J.A.L. Silva, J.J.R. Frau
´sto
da Silva, A.J.L. Pombeiro, J. Catal. 248 (2007) 130–136.
[31] G.S. Mishra, A.J.L. Pombeiro, J. Mol. Catal. A: Chem. 239 (2005)
96–102.
[32] Z. Otwinowski, W. Minor, in: C.W. CarterJr., R.M. Sweet (Eds.),
Processing of X-ray Diffraction Data Collected in Oscillation Mode.
Methods in Enzymology, Macromolecular Crystallography, Part A,
vol. 276, Academic Press, New York, 1997, pp. 307–326.
[33] G.M. Sheldrick, SHELXS-97, Program for Crystal Structure Deter-
mination, University of Go
¨ttingen, Germany, 1997.
[34] G.M. Sheldrick, SHELXTL, version 6.14-1, Bruker AXS, Inc.,
Madison, WI, 2005.
[35] G.M. Sheldrick, SHELXL-97, Program for Crystal Structure Refine-
ment, University of Go
¨ttingen, Germany, 1997.
1194 Y.Y. Karabach et al. / Journal of Inorganic Biochemistry 102 (2008) 1190–1194
... This attributes to TEAs coordinated with the zinc ion. 22,26,27 3.3. Anode and Electrolyte SEM. ...
Article
Full-text available
The zinc-air batteries (ZABs) are regarded as the most potential energy storage device for the next generation. However, the zinc anode passivation and hydrogen evolution reaction (HER) in alkaline electrolyte situations inhibit the zinc plate working efficiency, which needs to improve zinc solvation and better electrolyte strategy. In this work, we propose a design of new electrolyte by using a polydentate ligand to stabilize the zinc ion divorced from the zinc anode. The formation of the passivation film is suppressed greatly, compared to the traditional electrolyte. The characterization result presents that the quantity of the passivation film is reduced to nearly 33% of pure KOH result. Besides, triethanolamine (TEA) as an anionic surfactant inhibits the HER effect to improve the efficiency of the zinc anode. The discharging and recycling test indicates that the specific capacity of the battery with the effect of TEA is improved to nearly 85 mA h/cm2 compared to 0.21 mA h/cm2 in 0.5 mol/L KOH, which is 350 times the result of the blank group. The electrochemical analysis results also indicate that zinc anode self-corrosion is palliated. With density function theory, calculation results prove the new complex existence and structure in electrolytes by the data of the molecular orbital (highest occupied molecular orbital-lowest unoccupied molecular orbital). A new theory of multi-dentate ligand inhibiting passivation is elicited and provides a new direction for ZABs' electrolyte design.
... To detect and determine the quantity of tryptophan in reallife samples, many methods have been reported, including, electrochemical sensors [19][20][21][22][23][24][25][26][27], HPLC, spectrophotometry and capillary electrophoresis [28,29]. Pyromellitic acid (PMA) was chosen for the development of electropolymerized ITO electrode, where PMA acts as a ligand [30] for molecular framework, supramolecular hydrogels [31] and coordination polymers [32] due to the presence of four carboxylic acid substituted on the aromatic ring. In this study, we used the electropolymerized pyromellitic acid (PMA) on indium tin oxide (ITO) to selectively quantify tryptophan in real-life samples. ...
Article
Herein reporting a sensitive electrochemical strategy for the quantification of tryptophan on indium tin oxide electrode with an electropolymerized pyromellitic acid. In all perspectives, the characteristic features of the developed electrode were consistent with the requirements of a sensor. The morphological and functional identification of the modified electrode were done by scanning electron microscopy and infrared spectroscopy. The proposed electrode material has long-term stability and is easy to handle for real-life sample analysis. The fabricated electrode has a linear range of 10 to 300 μM with a limit of detection of 1.15 μM. The calculated sensitivity of the electrode was 1.3 μM/μA/cm2.
Article
The reactions of Cu(II), Co(II), and Zn(II) terephthalates with hydroxyalkylamines (tris(2-hydroxyethyl)amine, bis(2-hydroxyethyl)amine, tris(hydroxymethyl)aminomethane, and bis(2-hydroxyethyl)aminotris(hydroxymethyl)methane) are studied for the first time. The structures and properties of the synthesized complexes are studied by IR spectroscopy, electronic absorption spectroscopy, mass spectrometry, and elemental and thermal analyses. The structure of the binuclear heteroligand [Cu2(TEA)2(Tph)]n·H2O complex is studied by single-crystal X-ray diffraction (XRD) (CIF file CCDC no. 2224437).
Article
Cyclohexane (CHA), cyclohexene (CHE), and 1,4-cyclohexadiene (CHD) derivatives have a wide variety of applications and scope in synthetic and industrial organic chemistry. These cyclic hydrocarbons are found in a variety of natural products, including herbal plants and fruits as basic fragrance components. The extraction and isolation of these essential hydrocarbons are too complex and difficult to pursue. However, the reduction of benzene and its derivatives is an essential and effective approach for producing CHA, CHE, and CHD derivatives. To obtain CHA, CHE, and CHD derivatives, several types of reduction techniques for benzene and its derivatives have been developed over the years, including various forms of catalytic hydrogenation, Benkeser reduction, classic and many modified Birch reductions such as ammonia-free, metal-free photochemical, solvent-free, electrochemical and electride-mediated Birch reduction. In this study, the comparison of chemoselectivity, regioselectivity, reaction conditions, reduction efficiency, and environmentally acceptable approaches of these techniques used for benzene reduction are summarized.
Article
This work reports the synthesis of a new water-soluble tri-hetero metallic polymer formulated as [Cu0.152 Mn0.848 (μ-dipic)2{Na2 (μ-H2O)4}]n·nH2O (1) where dipic²⁻ = pyridine 2,6-dicarboxylato. Elemental analysis, fourier transformed-infrared (FT-IR) spectroscopy, thermogravimetric analyses (TGA), differential thermal analysis (DTA), vibration sample magnetometer (VSM), and single-crystal X-ray diffraction (SC-XRD) analysis were used for characterization of polymer 1. According to crystal structure analysis, polymer 1 belongs to a monoclinic system with space group C2/m, while cations of Cu²⁺ and Mn²⁺ occupy 15.2% and 84.8% of octahedral sites, respectively. The divalent ions of Mn²⁺ and Cu²⁺ show coordination number six, which have significantly distorted octahedral geometry. The as-synthesized polymer presents a 3D network extended by monomeric [Mn/Cu(dipic)2]2– units and polymeric{[Na2(μ-H2O)4]²⁺}n chains with extensive hydrogen bonding between crystal water molecules and adjacent layers. Polymer 1 was used as homogeneous catalyst for the mild oxidation of cyclohexane to the corresponding products (alcohol and ketone) in the presence of hydrogen peroxide as an environmentally friendly oxidant. To obtain the highest catalytic performance, various reaction parameters such as temperature, reaction time, the molar ratio of nitric acid to the catalyst, and the molar ratio of oxidant to the catalyst, were optimized. As a result, polymer 1 exhibited the maximum catalytic performance with a total product yield of 39.2% under optimized conditions. The mechanism of cyclohexane oxidation catalyzed by polymer 1 was also discussed.
Article
Paramagnetic complexes of hydrometallatranes with Co2+ and Ni2+ were studied by high-resolution 1H and 13C NMR spectroscopy. The paramagnetic shifts of the signals in the 1H NMR spectra by ∼100 ppm downfield with respect to their position in the spectra of uncoordinated ligands were observed. A similar upfield shift by ∼400 ppm was observed in the 13C NMR spectra. Analysis showed that paramagnetic shifts are caused by the contact hyperfine or electron-nucleus interactions between the unpaired electrons and the resonating nuclei of the ligand molecules. It was found that the preferred tridentate N,O,O-state of the ligand (triethanolamine) in aqueous solutions is accompanied by the intraligand exchange, in which two out of three oxygen centers compete for coordination. Triethanolamine was found to undergo substitution with 1-methylimidazole under mild conditions with the formation of complexes of the composition M(MIm)4X2 (M = Co, Ni; X = Cl, AcO). This causes a paramagnetic shift of the signals in the 13C NMR spectra as a result of the hyperfine interaction of the MIm nuclei with the unpaired spins of the coordinating ion.
Article
Full-text available
Publisher Summary X-ray data can be collected with zero-, one-, and two-dimensional detectors, zero-dimensional (single counter) being the simplest and two-dimensional the most efficient in terms of measuring diffracted X-rays in all directions. To analyze the single-crystal diffraction data collected with these detectors, several computer programs have been developed. Two-dimensional detectors and related software are now predominantly used to measure and integrate diffraction from single crystals of biological macromolecules. Macromolecular crystallography is an iterative process. To monitor the progress, the HKL package provides two tools: (1) statistics, both weighted (χ 2 ) and unweighted (R-merge), where the Bayesian reasoning and multicomponent error model helps obtain proper error estimates and (2) visualization of the process, which helps an operator to confirm that the process of data reduction, including the resulting statistics, is correct and allows the evaluation of the problems for which there are no good statistical criteria. Visualization also provides confidence that the point of diminishing returns in data collection and reduction has been reached. At that point, the effort should be directed to solving the structure. The methods presented in the chapter have been applied to solve a large variety of problems, from inorganic molecules with 5 A unit cell to rotavirus of 700 A diameters crystallized in 700 × 1000 × 1400 A cell.
Book
hemistry is the science about breaking and forming of bonds between atoms. One of the most important processes for organic chemistry is breaking bonds C–H, as well as C–C in various compounds, and primarily, in hydrocarbons. Among hydrocarbons, saturated hydrocarbons, alkanes (methane, ethane, propane, hexane etc. ), are especially attractive as substrates for chemical transformations. This is because, on the one hand, alkanes are the main constituents of oil and natural gas, and consequently are the principal feedstocks for chemical industry. On the other hand, these substances are known to be the less reactive organic compounds. Saturated hydrocarbons may be called the “noble gases of organic chemistry” and, if so, the first representative of their family – methane – may be compared with extremely inert helium. As in all comparisons, this parallel between noble gases and alkanes is not fully accurate. Indeed the transformations of alkanes, including methane, have been known for a long time. These reactions involve the interaction with molecular oxygen from air (burning – the main source of energy!), as well as some mutual interconversions of saturated and unsaturated hydrocarbons. However, all these transformations occur at elevated temperatures (higher than 300–500 °C) and are usually characterized by a lack of selectivity. The conversion of alkanes into carbon dioxide and water during burning is an extremely valuable process – but not from a chemist viewpoint.
Article
This review describes examples of remarkable acceleration of metal-catalyzed oxidation reactions by certain additives. In some cases, reactions proceed 2 or 10 times more rapidly in comparison with the process in the additive’s absence, in other cases, reactions become possible only in the presence of the additive. Varying ligands at the metal center or additives, one can not only dramatically improve yields of oxygenates but also control the selectivity of the reaction. Understanding mechanisms of the additive’s action is very important for search of new efficient catalysts and catalytic systems. Additives considered in the review can play roles of the ligands at metal ion or proton or electron transfer reagents and they mimic certain enzymes (the active center or its environment). Often the mechanism of the effect of additives on the reaction rate and the product yield is unknown, and the main aim of the review is to attract investigator’s attention in creating new efficient catalytic systems, which contain not only a metal ion but also a necessary “additive”.
Article
The main advances in the title area are reviewed with emphasis on catalytic alkane functionalization, both organometallic and bioinorganic. Current challenges are discussed. A tunable, selective hydrocarbon functionalization is still out of reach and resolving mechanistic problems has proved particularly challenging.
Article
The regiospecificity and stereoselectivity of alkane hydroxylation and alkene epoxidation by the particulate methane monooxygenase from Methylococcus capsulatus (Bath) was evaluated over a range of substrates. Oxidation products were identified by conventional GC analysis, and the stereoselectivity of oxidation was determined by a combination of chiral GC and HPLC methods, as well as 1H NMR analysis of the corresponding (R)-2-acetoxy-2-phenylethanoate ester derivatives in the case of alkanol products. Alkane hydroxylation was found to proceed favoring attack at the C-2 position in all cases, and the stereoselectivity for n-butane and n-pentane was characterized by an enantiomeric excess of 46% and 80%, respectively, with preference for the (R)-alcohol noted for both substrates. Epoxides were formed with smaller stereoselectivities. Together, the regio- and stereoselectivity results suggest that an equilibrium of competing substrate binding modes exists. A simple substrate-binding model that incorporates preferential C-2 oxidation with the observed stereoselectivity of alkane hydroxylation is proposed, and hypotheses for the general mechanism are suggested and discussed.
Article
The bis(maltolato)oxovanadium complexes [VO(ma)2], cis-[VO(OCH3)(ma)2] and [VO(py)(ma)2] have been covalently bonded to carbamated modified silica gel and these systems are shown to serve as effective heterogeneous catalysts for cyclohexane oxidation by molecular oxygen without any additive. The [VO(ma)2] catalyst gives the best results which are further promoted in the presence of 2-pyrazinecarboxylic acid which acts as a co-catalyst, while picolinic acid proved to be almost inactive. The reaction occurs under mild conditions (175 °C, 10 atm O2) forming two major products, cyclohexanol and cyclohexanone in a smaller amount, with a good selectivity. The TGA analysis of the catalyst shows that it is stable up to 273 °C and inductive couple plasma (ICP) indicates a limited metal loss after 20 h use of the catalyst up to 175 °C. The morphology of the catalyst was analyzed by SEM. Evidence is presented in favour of the involvement of a free-radical mechanism.Graphical abstractBis(maltolato)oxovanadium complexes have been covalently bonded to carbamated modified silica gel and shown to act as catalysts for the oxidation of cyclohexane by dioxygen, to cyclohexanol and cyclohexanone with a good activity and selectivity, without any additive and under relatively mild reactions. The effects of various factors were investigated towards the optimization of the reaction conditions and evidence is presented for a free-radical mechanism.