Você está na página 1de 45

Coordination Chemistry

International Journal of Inorganic Chemistry


Coordination Chemistry
International Journal of Inorganic Chemistry
Coordination Chemistry
Copyright 2011 Hindawi Publishing Corporation. All rights reserved.
This is a focus issue published in volume 2011 of International Journal of Inorganic Chemistry. All articles are open access articles
distributed under the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any
medium, provided the original work is properly cited.
Editorial Board
Christopher Allen, USA
Hakan Arslan, Turkey
Peter Baran, USA
Ivano Bertini, Italy
George Britovsek, UK
Alfonso Castin eiras, Spain
Stephen Colbran, Australia
William Connick, USA
Maochun Hong, China
Yining Huang, Canada
M. Ishaque Khan, USA
Karl Kirchner, Austria
George Koutsantonis, Australia
Abdessadek Lachgar, USA
Wolfgang Linert, Austria
James K. McCusker, USA
Norbert W. Mitzel, Germany
Rabindranath Mukherjee, India
Luis Oro, Spain
Alvaro J. Pardey, Venezuela
James E. Penner-Hahn, USA
Maurizio Peruzzini, Italy
Rainer Pottgen, Germany
Stephen Ralph, Australia
Daniel L. Reger, USA
Herbert W. Roesky, Germany
Axel Schulz, Germany
Konrad Seppelt, Germany
E. I. Solomon, USA
Alexander Steiner, UK
Wei-Yin Sun, China
W. T. Wong, Hong Kong
John Derek Woollins, UK
Contents
Syntheses and Crystal Structures of Two Transition Metal Complexes (M= Mn and Co) Containing
Malonate and Reduced Imino Nitroxide Radicals, Jing Chen, You-Juan Zhang, Bing-Chang Qin,
Hui-Min Zhu, and Yu Zhu
Volume 2011, Article ID 257521, 5 pages
Synthesis, Characterization, and Magnetic and Thermal Studies on Some Metal(II) Thiophenyl Schi
Base Complexes, Aderoju Amoke Osowole
Volume 2011, Article ID 650186, 7 pages
Synthesis and Structural Characterization of a New Tetranuclear Nickel(II) Sulfato Complex Containing
the Anionic Form of Di-2-Pyridyl Ketone Oxime, Eleni Moushi, Constantinos G. Efthymiou,
Spyros P. Perlepes, and Constantina Papatriantafyllopoulou
Volume 2011, Article ID 606271, 9 pages
Improvement of Aminopeptidase Activity of Dizinc(II) Complexes by Increasing Substrate Accessibility,
Md. Jamil Hossain, Akinobu Wada, Yasuhiro Igarashi, Kei-ichiro Aimono, Keisuke Suzuki, Katsuya Tone,
and Hiroshi Sakiyama
Volume 2011, Article ID 395418, 4 pages
Synthesis and Crystal Structure Dierences between Fully and Partially Fluorinated -Diketonate Metal
(Co
2+
, Ni
2+
, and Cu
2+
) Complexes, Akiko Hori and Masaya Mizutani
Volume 2011, Article ID 291567, 8 pages
A Selective Chemosensor for Mercuric Ions Based on 4-Aminothiophenol-Ruthenium(II)
Bis(bipyridine) Complex, Amer A. G. Al Abdel Hamid, Mohammad Al-Khateeb, Ziyad A. Tahat,
Mahmoud Qudah, Safwan M. Obeidat, and Abdel Monem Rawashdeh
Volume 2011, Article ID 843051, 6 pages
Hindawi Publishing Corporation
International Journal of Inorganic Chemistry
Volume 2011, Article ID 257521, 5 pages
doi:10.1155/2011/257521
Research Article
Syntheses and Crystal Structures of Two Transition Metal
Complexes (M= Mn and Co) Containing Malonate and Reduced
Imino Nitroxide Radicals
Jing Chen,
1
You-Juan Zhang,
1
Bing-Chang Qin,
1
Hui-Min Zhu,
1
and Yu Zhu
2
1
School of Chemistry and Chemical Engineering, Anyang Normal University, Anyang 455002, China
2
Department of Chemistry, Zhengzhou University, Zhengzhou 450052, China
Correspondence should be addressed to Jing Chen, chenjing64@eyou.com
Received 9 August 2010; Accepted 30 October 2010
Academic Editor: Rabindranath Mukherjee
Copyright 2011 Jing Chen et al. This is an open access article distributed under the Creative Commons Attribution License,
which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.
Two novel transition metal complexes with malonate and reduced imino nitroxide radicals, [Co(mal)(Him2-py)
2
] (ClO
4
)1 and
[Mn(mal)(Him2-py)
2
] (H
2
O)2 (Him2-py = 1-hydroxy-2-(2

-pyridyl)-4,4,5,5-tetramethylimidazoline) have been synthesized and


their crystal structures were determined by X-ray diraction method. During the reaction, one-electron reduction of the NO
radical moiety in IM2py has been reviewed. The structural analyses reveal that two title complexes are isostructural and crystallize
in monoclinic space group C2. For the complex 1, a = 17.004(9), b = 10.753(5), c = 9.207(5)

A with = 113.856(8)

. For the
complex 2, a = 16.721(5), b = 10.897(5), c = 9.253(3)

A with = 120.807(6)

. In two complexes, the coordination number


around the metal ion is six, and the coordination sphere is a distorted octahedron. Two nitrogen atoms from Him2-py and two
oxygen atoms from malonate are in the basal plane, and two nitrogen atoms from pyridyl rings of Him2-py at the axial position.
1. Introduction
The design and synthesis of transition metal complexes with
organic free radicals is one of the major challenges in the
eld of molecular magnetic materials [1]. Nitronyl nitroxide
(NN) radicals are normally used as spin carriers to the devel-
opment of molecular-based magnetic materials. However
nitroxides can undergo redox reactions with transition metal
ions under certain conditions [2, 3]. In fact, nitronyl free
radicals are in an oxidation state intermediate between those
of the hydroxylamino anion and the nitrosonium cation.
Up to now, relatively little work has been devoted to the
study of the redox properties of metal-nitroxyl systems and
only a few complexes containing metal ions bound to the
reduced monoradical have been reported [49]. It is known
that nitronyl nitroxide radicals can undergo redox reaction
with transition metal ions, yielding complexes in which the
IMHR reduced form of IM acts as a diamagnetic ligand
[7]. In order to extend our knowledge of extremely rich
chemistry of such systems, it is necessary to further explore
the reactions between metal ion and nitronyl nitroxide
radicals. In this paper, we will report that syntheses and
structural characterization about two novel transition metal
compounds with malonate and reduced imino nitroxide rad-
icals, [Co(mal)(Him2-py)
2
] (ClO
4
) 1 and [Mn (mal)(Him2-
py)
2
] (H
2
O) 2 (Him2-py = 1-hydroxy-2-(2

-pyridyl)-4,4,5,5-
tetramethylimidazoline).
2. Experimental
2.1. Syntheses. 2-(2

-pyridyl)-4,4,5,5-tetramethylimidazoli-
ne-1-oxyl (im2-py) was prepared according to the methods
reported [10].
[Co(mal)(Him2-py)
2
] (ClO
4
) 1: an aqueous solution
(10 mL) of Na
2
(mal) (0.148 g,1 mmol) was added to a
mixture of Co(ClO
4
)
2
6H
2
O (0.365 g,1 mmol) and im2-
py (0.436 g, 2 mmol) in 25 mL of methanol (pH = 6
8). The mixture was stirred for 2 h and ltered. The
ltrate was kept at room temperature for 1 month to
grow well-formed orange crystals of [Co(mal)(Him2-py)
2
]
(ClO
4
). Yield: 45%. Anal. Cacld. (%) for: C, 47.38; H,
4.96; N, 12.34. Found (%): C, 46.52; H, 4.92; N, 12.06. IR
(KBr):
(Py)
1448 cm
1
, 1386 cm
1
,
(NO)
1357 cm
1
,
ascoo
1650 cm
1
,
ascoo
1455 cm
1
.
2 International Journal of Inorganic Chemistry
Table 1: Summary of the crystallographic data and collections for two complexes.
Complex 1 Complex 2
Empirical formula C
27
H
34
ClCoN
6
O
10
C
27
H
36
MnN
6
O
7
Formula weight 696.98 611.56
Crystal system, space group Monoclinic, C2 Monoclinic, C2
Unit cell dimensions
a = 17.004(9)

A
b = 10.753(5)

A
c = 9.207(5)

A
= 13.856(8)

a = 16.721(5)

A
b = 10.897(5)

A
c = 9.253(3)

A
= 120.807(6)

Volume 1539.6(14)

A
3
1448.1(9)

A
3
Z, Calculated density 2, 1.504 Mg/m
3
2, 1.403 Mg/m
3
Absorption coecient 0.710 mm
1
0.511 mm
1
F(000) 724 642
Crystal size 0.18 0.16 0.14 mm
3
0.24 0.22 0.18 mm
3
range for data collection 2.30 to 25.10

2.35 to 25.01

Limiting indices
20 h 20, 12 k 11,
9 l 10
19 h 14, 12 k 12,
11 l 11
Reections collected/unique 3755/2253 (R(int) = 0.0318) 3730/2519 (R(int) = 0.0154)
Completeness to = 25.10 95.4% 100.0%
Absorption correction Semi-empirical from equivalents Semi-empirical from equivalents
Max. and min. transmission 1.000000 and 0.711602 1.000000 and 0.754606
Renement method Full-matrix least-squares on F
2
Full-matrix least-squares on F
2
Data/restraints/parameters 2253/47/228 2519/2/190
Goodness-of-t on F
2
1.037 1.107
Final R indices (I > 2(I)) R1 = 0.0502, wR2 = 0.1257 R1 = 0.0401, wR2 = 0.1069
R indices (all data) R1 = 0 .0583, wR2 = 0.1311 R1 = 0.0418, wR2 = 0.1091
Absolute structure parameter 0.00(3) 0.0(2)
Largest di. peak and hole 1.136 and 0.285 e.A
3
0.745 and 0.294 e.A
3
[Mn (mal)(Him2-py)
2
] (H
2
O) 2 was obtained using the
same procedure as that of the complex 1. The ltrate was
kept at room temperature for 20 days to grow well-formed
orange crystals of [Mn (mal)(Him2-py)
2
] (H
2
O). The yield
was about 55%. Anal. Cacld. (%) for: C, 53.12; H, 6.01; N,
14.02. Found (%): C, 52.72; H, 5.93; N, 13.74. IR (KBr):

(Py)
14 50 cm
1
, 1396 cm
1
,
(NO)
1365 cm
1
,
ascoo
1645
cm
1
,
scoo
1471 cm
1
.
2.2. Crystal Structure Determination and Renement. All
measurements were made on a Bruker Smart 1000 dirac-
tometer equipped with graphite-monochromated MoK
radiation ( = 0.71073

A). The data were collected at room
temperature. A summary of the crystallographic data is given
in Table 1. These structures were solved by direct methods
using the SHELXS97 program[11]. Full-matrix least-squares
renements on F
2
were carried out using SHELXL97 [12].
A summary of the crystallographic data and collections are
listed in Table 1. These signicant bond parameters for two
complexes are given in Tables 2 and 3, respectively. Views
of the molecular structures for compounds 1 and 2 are
shown, respectively, in Figures 1 and 2. The sketch of the
intermolecular hydrog bonds of two complexes are shown in
Figure 3 and 4, respectively.
3. Results and Discussion
The data of (=
as

s
) of IR reveal that each malonate
dianion binds metal ions in bidentate mode, leading to a
mononuclear structure. The bonds of the NO stretching
vibration appear at ca.1357 cm
1
and 1365 cm
1
for complex
1 and 2, respectively, which suggest that one Him2py are
chelated metal ion by pyridyl and imino nitrogen atoms with
the ve-membered ring.
The crystal structures of both complexes have several
features in common. The single-crystal X-ray structures
of complexes 1 (Figure 1) and 2 (Figure 2) conrm the
bidentate chelation of ligand. The complexes 1 and 2 consist
of mononuclear molecule [Co(mal)(Him2-py)
2
] (ClO
4
) and
[Mn (mal)(Him2-py)
2
] (H
2
O). The metal ion is located
International Journal of Inorganic Chemistry 3
Table 2: Selected bond distances (

A) and bond angles (

) for the
complex 1.
Co(1)O(2) 1.879(4) C(1)C(2) 1.360(9)
Co(1)N(1) 1.945(4) O(2)Co(1)O(2)#1 95.0(3)
Co(1)N(2) 1.947(5) O(2)Co(1)N(1)#1 93.8(2)
O(1)N(3) 1.400(6) O(2)Co(1)N(1) 86.8(2)
N(1)C(5) 1.336(7) N(1)#Co(1)N(1) 179.1(4)
N(1)C(1) 1.354(8) O(2)Co(1)N(2)#1 175.0(2)
N(2)C(6) 1.296(7) N(1)Co(1)N(2)#1 97.2(2)
N(2)C(10) 1.490(8) O(2)Co(1)N(2) 87.67(16)
N(3)C(6) 1.336(7) N(1)Co(1)N(2) 82.1(2)
N(3)C(7) 1.469(8) N(2)#1Co(1)N(2) 90.0(3)
Symmetry transformations used to generate equivalent atoms: no. 1
x, y, z.
O3
Co
O
N
C
H
Cl
C13
N1
N1
O1
N3
C6
N2
N3
N2
Co1
O2
O2
a
b
c
Figure 1: Diamond views of complex 1. Hydrogen atoms are
omitted for clarity.
in a distorted octahedral environment, formed by the four
nitrogen atoms (N(1), N(2), N(1)#, N(2)#) of the two
bidentate imino nitroxide radicals and two oxygen atoms
(O(2), O(2)#) of the same malonate group. The axial
positions are occupied by nitrogen atoms (N(1), N(1)#) from
pyridyl rings. The pyridyl rings of the im2-py ligands are
nonplanar, and that is a consequence of steric crowding from
the im2-py ligands of the same molecule.
For complex 1, the distances of Co(1)O(2) is 1.879(4)

A
and the NO bond distances are 1.400(6)

A, which show
that the complex is consisted of Co(III) and reduced species
IMHR [7]. For 2, the length of Mn(1)O(2) is 2.022(3)

A
and that of NO is 1.406(4)

A, clearly indicative of the
reduced form of the radical [1316]. In complex 1, the C2
N1 and C2N2 bond lengths of the IMH2py are 1.296(7)
Table 3: Selected bond distances (

A) and bond angles (

) for the
complex 2.
Mn(1)O(2) 2.022(3) N(3)C(7) 1.468(5)
Mn(1)N(2) 2.075(4) C(1)C(2) 1.379(7)
Mn(1)N(1) 2.116(3) O(2)#1Mn(1)O(2) 88.88(19)
O(1)N(3) 1.406(4) O(2)#1Mn(1)N(2) 167.69(15)
O(2)C(13) 1.251(6) O(2)Mn(1)N(2) 89.92(11)
O(3)C(13) 1.238(6) N(2)Mn(1)N(2)#1 93.8(2)
N(1)C(1) 1.312(5) N(2)Mn(1)N(1)#1 100.87(15)
N(1)C(5) 1.360(5) O(2)#1Mn(1)N(1) 90.13(15)
N(2)C(6) 1.287(5) O(2)Mn(1)N(1) 91.38(16)
N(2)C(10) 1.485(5) N(2)Mn(1)N(1) 77.65(15)
N(3)C(6) 1.367(5) N(1)#1Mn(1)N(1) 177.9(3)
Symmetry transformations used to generate equivalent atoms: no. 1
x, y, z.
O3
Mn
O
N
C
H
C13
N1
N1
C5
C5
Mn1
N3
C6
C6
N2
N3
N2
O1
O1
O2
O2
a
b
c
Figure 2: Diamond views of complex 2. Hydrogen atoms are
omitted for clarity.
and 1.336(7)

A, respectively, as well as the C2N1 and C2
N2 bond lengths of the IMH2py are 1.287(5) and 1.367
(5)

A in complex 2, which are rather close to those of the
reduced imino nitroxide complex [69]. These structural
changes result from the one-electron reduction of the N
O radical moiety in IM2py. N(1)MN(1)# angles do
not signicantly deviate from orthogonal, ranging from
4 International Journal of Inorganic Chemistry
Figure 3: View of the unit of the complex 1 showing 1-D zigzag chain formed by OO interaction alone a axis.
Figure 4: View of the unit of the complex 2, showing 1-D zigzag chain formed by OO interaction alone a axis.
179.1(4)

to 177.9(3)

. Details of two complexes can be


found in supplementary material available online at doi:
10.1155/2011/257521. The dihedral angles between pyridyl
ring (N1C5C4C1) and imino nitroxide group (N2C6N3O1)
are 12.8

(for complex 1) and 12.9

(for complex 2),


respectively. For complex 1, noncoordinated ClO

4
anions
insert in the crystal spacing as well as noncoordinated H
2
O
molecules insert there for complex 2.
Sketch of the intermolecular hydrogen bonds of the
complex 1 and 2 are shown in Figure 2, 3, respectively.
In the crystal packing of two complexes, hydrogen bonds
of OO type have been observed between the hydrogen
atom from the coordinated malonate and the oxygen atom
from adjacent Him2-py, thus one-dimensional structures are
formed. Among the 1-D chains, the oxygen atoms from the
NO groups of Him-2py and the mal form hydrogen bonds,
and the distances of O(+x, +y, +z)O(1/2 x, 1/2 + y, z)
are 2.612

A (for complex 1) and 2.627

A (for complex 2),
respectively.
As we know, there were several reports on metal
complexes with a reduced nitroxide radical. However, no
example in which M(mal)
2
converts the IMR radical into
IMHR has been reported. The mechanismic details of the
reduction of nitroxide radical are not completely clear, but
it is likely that the formation of Him2-py is favored by acidic
impurities and standing for a long time. The reduced radical
can exist in two tautomeric forms [7], the amidino oxide
and iminohydroxylamine. Since the nitrogen atom from the
imino group here is involved in the coordinating metal ion,
the reduced radical form should be the latter [6].
International Journal of Inorganic Chemistry 5
As mentioned in the experimental part of this paper,
the cobalt reactant is cobalt(II) perchlorate hexahydrate.
However, the reaction product is cobalt(III) complex 1 due
to the oxidation of Co(II) to Co(III) in air. The result is
corresponding to those lectures [1719].
Acknowledgment
This work was supported by the Program for New Century
Excellent Talents in University of Henan Province (no.
2006HANCET-14).
References
[1] D. Luneau and P. Rey, Magnetism of metal-nitroxide com-
pounds involving bis-chelating imidazole and benzimidazole
substituted nitronyl nitroxide free radicals, Coordination
Chemistry Reviews, vol. 249, no. 23, pp. 25912611, 2005.
[2] M. H. Dickman and R. J. Doedens, Structure of chloro
(2,2,4,4-tetramethylpiperidinyl-1-oxo-O,N) (triphenylphosp-
hine) palladium(II), a metal complex of a reduced nitroxyl
radical, Inorganic Chemistry, vol. 21, no. 2, pp. 682684, 1982.
[3] L. C. Porter and R. J. Doedens, (Hexauoroacetylacetonato-
O, O

)(1-hydroxy-2,2,6,6-tetramethylpiperidinato-
O,N)palladium(II), [Pd(C
5
HF
6
O
2
)(C
9
H
18
NO)], a metal
complex containing a reduced nitroxyl radical, Acta
Crystallographica Section C, vol. 41, no. 6, pp. 838840, 1985.
[4] A. Caneschi, D. Gatteschi, J. Laugier, P. Rey, and C. Zanchini,
Synthesis, X-ray crystal structure, and magnetic properties of
two dinuclear manganese(II) compounds containing nitronyl
nitroxides, imino nitroxides, and their reduced derivatives,
Inorganic Chemistry, vol. 28, no. 10, pp. 19691975, 1989.
[5] A. Caneschi, D. Gatteschi, M. C. Melandri, P. Rey, and R.
Sessoli, Structure and magnetic properties of manganese(II)
carboxylate chains with nitronyl nitroxides and their reduced
amidino-oxide derivatives. From random-exchange one-
dimensional to two-dimensional magnetic materials, Inor-
ganic Chemistry, vol. 29, no. 21, pp. 42284234, 1990.
[6] L. Li, D. Liao, L. Bai, Z. Jiang, and S. Yan, Synthesis and
crystal structure of oxalato-bridged dicopper(II) complex
with reduced imino nitroxide radicals, Journal of Molecular
Structure, vol. 569, no. 13, pp. 179183, 2001.
[7] T. Ishida, T. Suzuki, and S. Kaizaki, Synthesis and charac-
terization of [Co(NO
2
)
2
(acac) (IMH2py)] with one-electron-
reduced imino nitroxide radical associated with unusual
displacement of acetylacetonate in the starting complex trans-
Na[Co(NO
2
)
2
(acac)
2
], Inorganica Chimica Acta, vol. 357, no.
11, pp. 31343138, 2004.
[8] QI. H. Zhao, L. C. Li, Z. H. Jiang, D. Z. Liao, S. P. Yan, and R. B.
Fang, Synthesis and crystal structure of a new copper(II) bin-
uclear complex bridged by the reduced derivative of a nitronyl
nitroxide biradical, Journal of Coordination Chemistry, vol. 57,
no. 10, pp. 843848, 2004.
[9] L.-Y. Wang, C.-X. Zhang, D.-Z. Liao, Z.-H. Jiang, and S.-
P. Yan, Synthesis and crystal structure of an azide bridged
binuclear zinc(II) complex including the reduced derivative of
nitronyl nitroxide, [Zn(Him2Py)(N
3
)
2
]
2
, Chinese Journal of
Structural Chemistry, vol. 23, no. 2, pp. 171175, 2004.
[10] E. F. Ullman, L. Call, and J. H. Osiecki, Stable free radicals.
VIII. New imino, amidino, and carbamoyl nitroxides, Journal
of Organic Chemistry, vol. 35, no. 11, pp. 36233631, 1970.
[11] G. M. Sheldrick, SHELXS97, Program for the Solution of
CrystalStructure, University of Gottingen, Germany, 1997.
[12] G. M. Sheldrick, SHELXL97, Program for the Renement of
Crystal Structure, University of Gottingen, Germany, 1997.
[13] O. P. Anderson and T. C. Kuechler, Crystal and molec-
ular structure of a nitroxyl radical complex of cop-
per (II): Bis(hexauoroacetylacetonato)(4-hydroxy-2,2,6,6-
tetramethylpiperidinyl-N-oxy) copper(II), Inorganic Chem-
istry, vol. 19, no. 6, pp. 14171422, 1980.
[14] C. Benelli, D. Gatteschi, C. Zanchini, R. J. Doedens, M. H.
Dickman, and L. C. Porter, EPR spectra of bis(nitroxyl)
adducts of bis(hexauoroacetylacetonato)manganese(II),
Inorganic Chemistry, vol. 25, no. 19, pp. 34533457, 1986.
[15] H. Oshio, T. Watanabe, A. Ohto, T. Ito, and H. Masuda,
Intermolecular Ferromagnetic and Antiferromagnetic Inter-
actions in Halogen-Bridged Copper(I) Imino Nitroxides: crys-
tal Structures and Magnetic Properties of [Cu
I
(-X)(imino
nitroxide)]
2
(X = I or Br), Inorganic Chemistry, vol. 35, no.
2, pp. 472479, 1996.
[16] M. Fettouhi, M. Khaled, A. Waheed et al., Manganese (II)
coordination complexes involving nitronyl nitroxide radicals,
Inorganic Chemistry, vol. 38, no. 18, pp. 39673971, 1999.
[17] S. Ross, T. Weyherm uller, E. Bill, K. Wieghardt, and P. Chaud-
huri, Tris(pyridinealdoximato)metal complexes as ligands for
the synthesis of asymmetric heterodinuclear Cr
III
M species
[M = Zn(II), Cu(II), Ni(II), Fe(II), Mn(II), Cr(II), Co(III)]:
a magneto-structural study, Inorganic Chemistry, vol. 40, no.
26, pp. 66566665, 2001.
[18] Y. L. Feng and S. X. Liu, Spectral Property and Crystal Struc-
ture of Cobalt (III) Complexes of Isonitrosoacetylacetone-N-
arylimine, Yingyong Huaxue, vol. 19, no. 9, p. 842, 2002.
[19] Z.-J. Xiao and S.-X. Liu, Synthesis and extended 3D
hydrogen-bonded structure of complex tris(violurato-N,O)
cobalt(III) hexahydrate, Chinese Journal of Structural Chem-
istry, vol. 25, no. 2, pp. 163167, 2006.
Hindawi Publishing Corporation
International Journal of Inorganic Chemistry
Volume 2011, Article ID 650186, 7 pages
doi:10.1155/2011/650186
Research Article
Synthesis, Characterization, and Magnetic and Thermal Studies
on Some Metal(II) Thiophenyl Schiff Base Complexes
Aderoju Amoke Osowole
Inorganic Chemistry Unit, Department of Chemistry, University of Ibadan, Ibadan, Nigeria
Correspondence should be addressed to Aderoju Amoke Osowole, aderoju30@yahoo.com
Received 7 November 2010; Revised 22 December 2010; Accepted 31 December 2010
Academic Editor: Maurizio Peruzzini
Copyright 2011 Aderoju Amoke Osowole. This is an open access article distributed under the Creative Commons Attribution
License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly
cited.
4-(Thiophen-3-yl)-aniline undergoes condensation with o-vanillin to form an ONS donor Schi base, 2-methoxy-6-[(4-
thiophene-3-yl-phenylimino)-methyl]-phenol, which forms complexes of the type [ML
2
]xH
2
O (where M = Mn, Co, Ni, Cu, Zn,
Pd). These complexes are characterized by elemental analysis,
1
H nmr, electronic, mass, and IR spectroscopies and conductance
measurements. The electronic, IR and CHN data are supportive of a 4-coordinate tetrahedral geometry for Mn(II), Co(II), Ni(II),
and Zn(II) complexes and square-planar geometry for Cu(II) and Pd(II) complexes, with the chromophores N
2
O
2
. The magnetic
data reveals that the complexes are magnetically dilute and mononuclear with exception of the Cu(II) complex, which exhibits
some anti-ferromagnetisms. The complexes are air-stable solids, and none is an electrolyte in nitro methane.
1. Introduction
My group has in the last ve years been actively involved in
the synthesis, characterization, and thermal and biological
properties of various Schi bases and their M(II) chelates
(M = VO, Mn, Co, Ni, Cu, Zn, Pd), with the objectives of
deriving Schi base chelates which can be used as precursors
in metal-organic chemical vapor depositions (MOCVD) and
those with good in vitro antimicrobial activities as surface
cleaning agents [15]. Thiophenyl Schi bases are particu-
larly interesting because of their wide range of activities such
as anticancer activity as shown by benzyl-N-[1-(thiophenyl-
3-yl)ethylidene] hydrazine carbodithioate [6], antibacterial
and antifungal activities typied by thiophenyl-azetidinones,
-cephalexins and -vinyl anilines [716] as well as optical
exhibited by thiophene-2-aldazine [17]. Other activities
include structural activity exemplied by the hexacar-
bonyldiiron complex of N-(2-thienylmethylidene)aniline,
which is composed of two thienyl moieties derived from
the original thienyl imine and coupled together by a CC
bond [18]. Furthermore, Palladium imine complexes of 2-
thiophenecarboxaldehyde are used as catalysts in the Suzuki
cross-coupling of aryl bromides with phenyl boric acid
[19]. Extensive literature reviews show that no studies are
reported on the Schi base derived from o-vanillin and 4-
(thiophene-3-yl)-aniline and its metal(II) chelates [619].
Thus, the objective of this work is to synthesize, characterize,
and investigate the magnetic and thermal properties of the
Schi base, 2-methoxy-6-[(4-thiophene-3-yl-phenylimino)-
methyl]-phenol and its Mn(II), Co(II), Ni(II), Cu(II), Zn(II)
and Pd(II) complexes. These metal complexes and its ligands
are new, being reported here for the rst time.
2. Experimental Details
2.1. Materials and Physical Measurements. Reagent grade o-
vanillin, 4-(thiophene-3-yl)-aniline, manganese(II) nitrate
dehydrate, cobalt(II) nitrate hexahydrate, nickel(II) nitrate
hexahydrate, copper(II) nitrate hexahydrate, zinc(II) nitrate
hexahydrate, and palladium(II) chloride are purchased from
BDH and Aldrich chemicals and are used as received. Sol-
vents are dried and distilled before use according to standard
procedures. Melting points (uncorrected) are determined
using the Stuart scientic melting point SMP1 machine,
and conductivities of 10
3
M solutions of the complexes are
measured in nitromethane at 25

C using a MC-1, Mark V


conductivity meter with a cell constant of 1.0. The solid
2 International Journal of Inorganic Chemistry
reectance spectra are recorded on a Perkin-Elmer 20 spec-
trophotometer while infrared spectra are measured as KBr
discs on a Perkin-Elmer FTIR paragon 1000 spectrometer in
the range 4000400 cm
1
. The elemental analyses C, H, and
N are recorded on GmbH VarioEl analyser, and manganese,
cobalt, nickel, copper, zinc, and palladium are determined
titrimetrically and by atomic absorption spectroscopy [20].
The
1
H nmr spectra are recorded on a 300 MHz Oxford
Varian NMR instrument in CDCl
3
at 295 K.
1
H chemical
shifts are referenced to the residual signals of the protons
of CDCl
3
and are quoted in ppm. Magnetic susceptibilities
are measured on Johnson Matthey magnetic susceptibility
balance, and diamagnetic corrections are calculated using
Pascals constants [21]. Thermogravimetric analyses are done
in static air, using a T6 Linseis thermal analyser with a heat-
ing rate of 10

C/min in the range 30700

C. The MALDI-
TOF mass and atomic absorption spectra are obtained using
a Bruker Daltonic Reex TOF spectrometer with graphite as
matrix and Perkin Elmer Analyst 200 coupled to Winlab 32
software assembly, respectively.
2.2. Preparation of the Schi Base (2-methoxy-6-[(4-thio-
phene-3-yl-phenylimino)-methyl]-phenol). A 20 mL solution
of 8.71 mmol (1.33 g) o-vanillin in absolute ethanol is added
dropwise to a stirring solution of 8.71 mmol (1.53 g) of 4-
(thiophene-3-yl)-aniline in 30 mL of absolute ethanol. The
resulting orange-colored solution is reuxed for 4 h after
addition of 4 drops of acetic acid. The orange product
formed on cooling to room temperature is ltered and
recrystallized from ethanol. The yield of the title compound
is 1.88 g (70%).
1
H nmr (ppm) 11.13 (s, 1H, OH), 9.93
(s, 1H, HCN), 8.70 (1H, s, C
2
HS), 7.357.43 (m, 3H, C
6
H
3
);
7.037.24 (m, 4H, C
6
H
4
); 6.917.01 (m, 2H, C
2
H
2
S); 3.96 (s,
3H, OCH
3
).
2.3. Preparation of the Metal(II) Complexes. The various
complexes are prepared by gradual addition of 0.35 mmol
(0.060.11 g) M(NO
3
)
2
6H
2
O (M = Mn, Co, Ni, Cu, Zn)
neat to a stirring 0.7 mmol (0.22 g) of the ligand in 30 mL
of absolute ethanol. The resulting solutions are then buered
with 0.7 mmol (0.10 mL) of triethylamine and reuxed for
6 h during which the products formed. The precipitated
solids are ltered, washed with ethanol, and dried over
anhydrous calcium chloride. The yields are 0.12 g (50%),
0.17 g (70%), 0.17 g (70%), 0.17 g (70%), and 0.14 g (60%),
respectively.
The Pd(II) complex is prepared using a similar method.
0.36 mmol (0.064 g) of Pd(II) chloride, in 10 mL of abso-
lute ethanol, is added dropwise to a stirring solution
of 0.72 mmol (0.23 g) of the ligand in 30 mL of abso-
lute ethanol. The resulting solution is then buered with
0.72 mmol (0.11 mL) of triethylamine and reuxed for 6 h
during which the product is formed. The product is ltered,
washed with ethanol, and dried over anhydrous calcium
chloride. The yield is 0.18 g (70%).
Proton nmr measurements are done only for the diamag-
netic Zn(II) and Pd(II) complexes. Zn(II) complex:
1
H nmr
(ppm) 8.70 (s, 1H, HCN), 8.34 (1H, s, C
2
HS), 7.287.43
(m, 3H, C
6
H
3
); 7.017.10 (m, 4H, C
6
H
4
); 6.636.94 (m, 2H,
C
2
H
2
S); 3.94 (s, 3H, OCH
3
).
Pd(II) complex:
1
H nmr (ppm) 8.67 (s, 1H, HCN), 8.30
(1H, s, C
2
HS), 7.487.66 (m, 3H, C
6
H
3
); 7.007.47 (m, 4H,
C
6
H
4
); 6.886.98 (m, 2H, C
2
H
2
S); 3.94 (s, 3H, OCH
3
).
3. Results and Discussion
The equations for the formation of the complexes are
M(NO
3
)
2
+ 2HL [ML
2
] + 2HNO
3
(1)
(where M = Zn(II), Cu(II), Mn(II), Co(II), Ni(II))
PdCl
2
+ 2HL [PdL
2
] + 2HCl. (2)
All complexes adopt [ML
2
] stoichiometry, except Mn(II),
Co(II) and Ni(II) complexes that formas [ML
2
]xH
2
O, where
x = 1 and 0.5, respectively. Proposed structures for the
ligand and the Cu(II) complex are shown in Figure 1. The
formation of this ligand is conrmed by microanalysis and
1
H nmr. The colors, melting points, and room temperature
magnetic moments (
e
) of the compounds are presented in
Table 1. Attempts to isolate suitable crystals for single X-ray
structural determination have not been successful so far.
3.1. Infrared Spectra. The relevant infrared bands of the
compounds are presented in Table 2. The broadband at
3360 cm
1
in the ligand, which is conspicuously absent
in the spectra of the metal(II) Schi base complexes, is
assigned as vOH stretching frequency, and it conrms
involvement of the phenol O in chelation. It is broad due
to intramolecular hydrogen bonding, usually very strong
in Schi bases [1]. The new broadband at 3500 cm
1
in
the spectra of Co(II), Ni(II), and Mn(II) complexes is
assigned to the vOH frequency of crystallization H
2
O. The
uncoordinated C=N stretching vibrations are observed as
three bands between 16141521 cm
1
in the ligand [25]
and 16391503 cm
1
in the metal complexes with exceptions
of the Cu(II) and Pd(II) complex which have two bands.
The bathochromic/hypsochromic shifts of these bands in
the complexes are attributed to the involvement of N atom
of C=N in coordination to the metal ions. Moreover, it
has been documented that square planar Pd(II) and Cu(II)
Schi base complexes do exhibit geometric isomerism [22],
with the trans isomer showing two vC=N bands and the cis
isomer a lone vC=N band. The spectra of the Cu(II) and
Pd(II) complexes in this work show two vC=N bands and
are consequently in the trans-isomeric form. The vPh/CO
and CHvibrations of the ligand are observed at 14611364
and 972 cm
1
, respectively. These suer bathochromic shifts
to 12981177 and 896720 cm
1
in the Schi base complexes
due to the coordination of the phenol oxgen atom and
pseudoaromatic nature of the chelates [5, 6]. The observation
of new bands at 480405 and 581542 cm
1
due to v(MO)
and v(MN) [11, 14, 22] is further evidence of coordination.
3.2. Electronic Spectra and Magnetic Moments. The electronic
spectral data for the complexes are presented in Table 2.
International Journal of Inorganic Chemistry 3
OH
OCH
3
N
S
(a)
O
OCH
3
N
S
O
OCH
3
N
S
Cu
O
OCH
3
N
S
O
OCH
3
N
S
Cu
(b)
Figure 1: Proposed structures for the ligand (a) and its Cu(II) complex (b).
Table 1: Analytical data for the ligand and its complexes.
Compound
(empirical formula)
Formula
mass
Color
m/z
(100%)

e
% Yield
m

M.p
(

C)
Analysis (calculated)
% C % H % N % M
HL (C
18
H
15
NO
2
S) 309.38 Orange 309 80 150151
69.82
(69.88)
4.53
(4.89)
4.35
(4.53)

[MnL
2
]H
2
O
(MnC
36
H
30
N
2
S
2
O
5
)
689.73 green 671 5.70 50 10.0 338340
62.54
(62.69)
4.32
(4.38)
4.02
(4.06)
8.02
(7.97)
[CoL
2
]H
2
O
(CoC
36
H
30
N
2
S
2
O
5
)
693.71 Brick Red 675 4.33 70 20.0 290292
62.36
(62.33)
3.64
(4.36)
4.06
(4.04)
8.48
(8.50)
[NiL
2
]1/2H
2
O
(NiC
36
H
30
N
2
S
2
O
4.5
)
684.48 Yellow 675 3.10 70 15.0 193195
63.07
(63.17)
3.86
(4.27)
4.17
(4.09)
8.55
(8.58)
[CuL
2
]
2
(Cu
2
C
72
H
56
N
4
S
4
O
8
)
680.31 Brown 680 1.56 70 9.0 205207
64.14
(63.56)
4.06
(4.15)
4.01
(4.12)
9.35
(9.34)
[ZnL
2
]
(ZnC
36
H
28
N
2
S
2
O
4
)
681.76 Yellow 681 D 60 17.0 316318
63.13
(63.42)
3.84
(4.14)
3.81
(4.11)
9.52
(9.53)
[PdL
2
]
(PdC
36
H
28
N
2
S
2
O
4
)
723.18 green 720 D 70 12.0 148150
60.37
(59.79)
4.32
(3.90)
3.86
(3.87)
14.70
(14.72)

1
cm
2
mol
1
, D: diamagnetic.
The Mn(II) Schi base complex shows two absorption bands
at 15.00 and 23.11 kK, respectively, consistent with a four-
coordinate, tetrahedral geometry and are assigned to
6
A
1

4
E
1
(
1
) and
6
A
1

4
A
1
(
2
) transitions [23]. A room
temperature moment of 5.92 B.M is usually observed for the
Mn(II) compounds, regardless of stereochemistry because
the ground term is an
6
A
1
, and thus, orbital contribution
is nil. This Mn(II) complex has a moment of 5.70 B.M.
complementary of tetrahedral geometry [1].
The cobalt(II) Schi base complex gives two absorption
bands at 10.82 and 18.20 kK, respectively, typical of a four-
coordinate tetrahedral geometry and is assigned to
4
A
2

4
T
1
(P) (v
2
) and
4
A
2

4
T
1
(P) (v
3
). The transition
4
A
2

4
T
2
(v
1
) in the range 57 kK is not observed as usual
4 International Journal of Inorganic Chemistry
Table 2: Relevant infrared and electronic spectral data of the ligand and its complexes.
Compound OH (C=N) + (C=C) Ph/CO CH (MN) (MO) Electronic transitions (kK)
HL 3360b 1614s 1593s 1521s 1461s 1364s 972s 33.50, 36.25, 40.20
[MnL
2
]H
2
O 3500b 1612s 1589s 1551s 1298s 1194s 848s 720s 580m 561s 450m 420s 15.00, 23.11, 30.26, 33.45, 40.20
[CoL
2
]H
2
O 3500b 1609s 1557s 1503s 1231m 1180m 840s 777m 575s 537m 475m 415m 10.82, 18.20, 32.0, 37.40, 41.0
[NiL
2
]1/2H
2
O 3500b 1599s 1585s 1537s 1233s 1192s 853s 780m 581s 564s 460s 453s 14.29, 20.21, 30.72, 36.25, 40.40
[CuL
2
] 1587s 1538s 1237m 1193m 856s 795s 573s 557s 465s 410s 14.09, 21.60, 30.92, 34.35, 41.20
[ZnL
2
] 1609s 1582s 1504s 1231m 1197m 896s 731m 579s 562m 480m 405m 28.20, 35.00, 42.00
[PdL
2
] 1592s 1540s 1240s 1177s 843m 783m 562m 542m 450m 420m 18.20, 26.32, 30.25, 39.20, 43.00
b: broad, m: medium, s: strong., 1 kK: 1000 cm
1
.
since it lies in the infrared region [10]. This geometry is
corroborated by a moment of 4.33 B.M [11].
Nickel(II) complexes are known to exhibit complicated
equilibria between coordination numbers six (octahedral) to
four (square planar/tetrahedral) [24]. The Ni(II) Schi base
complex exhibits two absorption bands at 14.29 and 20.21 kK
typical of a 4-coordinate tetrahedral geometry, assigned
to
3
T
1
(F)
3
T
2
, (
2
) and
3
T
1
(F)
3
A
2
, (
3
) transitions.
Its moment of 3.10 B.M is complimentary of tetrahedral
geometry, since moments of 3.13.5 B.M. are reported for
distorted tetrahedral complexes [12].
The copper(II) complex displays two bands at 14.09
and 21.60 kK, assigned to
2
B
1g

2
A
1g
and
2
B
1g

2
E
1g
transitions of 4-coordinate, square planar geometry [13]. A
moment of 1.92.2 B.M. is usually observed for mononuclear
copper(II) complexes, regardless of stereochemistry [14]. A
magnetic moment of 1.56 B.M. is observed for this com-
plex, indicative of the presence of some anti-ferromagnetic
interactions, operating through CuCu interactions [25].
However, this could not be probed further due to lack of
facilities for variable temperature magnetic measurements
and nonsuitable crystal for single X-ray diraction measure-
ment (Figure 1).
The Zn(II) complex expectedly shows only charge trans-
fer transition from ML and -

transitions, as no
d-d transition is expected at 28.20, 35.00, and 42.0 kK,
respectively. This complex is diamagnetic, conrming its
tetrahedral geometry [1214].
The Pd(II) complex shows absorption bands at 18.20 and
26.32 kK, typical of square planar geometry and is assigned to
1
A
1g

1
B
1g
and
1
A
1g

1
E
2g
transitions. This complex is
expectedly diamagnetic [26].
3.3.
1
Hnmr Spectra. The ligand shows the phenolic proton as
a singlet at 11.13 ppm(s, 1H, OH), the imine (s, 1H, HCN),
and 2-thiophenyl (s, H, C
2
HS) protons resonate as singlets
at 9.93 and 8.70 ppm, respectively. The o-vanillin (m, 3H,
C
6
H
3
) and phenyl (aniline) protons (m, 4H, C
6
H
4
) are both
observed as multiplets at 7.357.43 and 7.037.24 ppm,
respectively. The two protons of thiophenyl ring at 4 and 5
positions come up as a multiplet at 6.917.01 ppm (m, 2H,
C
2
H
2
S), and the methoxy protons are observed as a singlet
at 3.96 ppm (s, 3H, OCH
3
).
In the Zn(II) complex spectrum, the phenolic proton
disappears, an indication of coordination of the phenol
oxygen to the Zn(II) ion, and other protons are all upeld
shifted in comparison to the ligand. The imine hydrogen
resonates as a singlet at 8.70 ppm (s, 1H, HCN) while the
2-thiophenyl proton is seen at 8.34 ppm as a singlet (s, H,
C
2
HS).
The o-vanillin protons resonate as a multiplet at 7.28
7.43 ppm (m, 3H, C
6
H
3
), and the phenyl(aniline) protons
are seen as multiplets at 7.017.10 ppm (m, 4H, C
6
H
4
).
The two protons of thiophenyl ring at 4 and 5 positions
resonate as a multiplet at 6.636.94 ppm (m, 2H, C
2
H
2
S),
and the methoxy protons are seen as a singlet at 3.94 ppm (s,
3H, OCH
3
). This shift shows deshielding, a consequence of
coordination of the imine nitrogen atom [27].
The Pd(II) complex spectrum reveals the absence of the
phenolic proton. This is indicative of involvement of the
phenol oxygen in coordination to the Pd(II) ion. The other
protons are all upeld shifted in comparison to the ligand
with exceptions of the o-vanillin and phenyl(aniline) protons
which are downeld shifted. The imine hydrogen (s, 1H,
HCN) and 2-thiophenyl (s, H, C
2
HS) protons resonate as
singlets at 8.67 and 8.30 ppm, respectively. The o-vanillin (m,
3H, C
6
H
3
), phenyl(aniline) (m, 4H, C
6
H
4
), and two protons
of thiophenyl ring (m, 2H, C
2
H
2
S) at 4 and 5 positions are
all seen as multiplets at 7.487.66, 7.007.47, and 6.88
6.98 ppm, respectively. The methoxy protons resonate as a
singlet at 3.94 ppm (s, 3H, OCH
3
). These shifts are indicative
of coordination of the imine nitrogen atom to the Pd(II) ion
[26].
3.4. Mass Spectroscopy and Thermal Studies. The mass spec-
tra of ligand and the complexes showed peaks attributed to
the molecular ions m/z at 309 [L]
+
; 671 [MnL
2
-2H]
+
; 675
[CoL
2
-2H]
+
; 675 [NiL
2
-2H]
+
; 680 [CuL
2
-2H]
+
; 681 [ZnL
2
-
2H]
+
and 720 [PdL
2
], respectively, and are presented in
Table 1.
The thermal degradation of the ligand and complexes
is presented in Table 3. The ligand, HL, decomposes in
three steps. First, the loss of the fragment C
2
H
2
and
0.5 mol N
2
at 30220

C, with mass losses of (obs. = 12.96%,


calc. = 12.93%). The next step involves the loss of the organic
fraction, C
11
H
8
O
2
S, with mass losses of (obs. = 66.32%,
International Journal of Inorganic Chemistry 5
Table 3: Thermal data for the ligand and its complexes.
Compound Temperature range (

C)
TG weight loss (%)
Assignments
Calc. Found
HL (C
18
H
15
NSO
2
)
30220 12.93 12.96 C
2
H
2
+ 0.5N
2
220420 65.94 66.32 C
11
H
8
O
2
S
420700 22.78 21.01 C
5
H
5
[MnL
2
]H
2
O (MnC
36
H
30
N
2
S
2
O
5
)
30200 9.57 9.40 H
2
O + 1.5O
2
210450 44.07 44.45 C
20
H
18
NS
450700 22.04 22.53 C
8
H
10
OS
Mn (residue)
[CoL
2
]H
2
O (CoC
36
H
30
N
2
S
2
O
5
)
30200 4.90 4.77 CH
4
+ H
2
O
200400 20.76 21.45 C
6
H
8
S
2
400700 57.08 57.50 C
24
H
16
O
4
N
2
Co (residue)
[NiL
2
]1/2H
2
O (NiC
36
H
30
N
2
S
2
O
4.5
)
30240 10.67 10.56 0.5H
2
O + SO
2
240420 23.08 22.73 C
10
H
10
N
2
420700 55.81 55.51 C
25
H
18
SO
2
Ni (residue)
[CuL
2
] (CuC
36
H
28
N
2
S
2
O
4
)
30200 17.93 17.86 C
8
H
12
N
200400 33.81 33.63 C
13
H
10
O
2
S
400700 38.81 42.00 C
15
H
6
NO
2
S
Cu (residue)
[ZnL
2
] (ZnC
36
H
28
N
2
S
2
O
4
)
30260 5.00 5.17 H
2
S
260440 27.87 27.80 C
11
H
10
OS
440700 52.22 52.21 C
22
H
16
N
2
O
3
Zn (residue)
[PdL
2
] (PdC
36
H
28
N
2
S
2
O
4
)
30210 3.87 3.50 C
2
H
4
210400 40.79 41.33 C
18
H
15
SO
2
400700 33.32 33.42 C
13
H
9
N
2
SO
PdO (residue)
calc. = 65.94%) at 220420

C. The nal step is the loss of


the fragment C
5
H
5
, with mass losses of (obs. = 22.78%,
calc. = 21.01%) at 420700

C.
The Mn(II) complex decomposes in three phases. The
rst phase corresponds to the loss of 1.5 moles of O
2
and
H
2
O between 30200

C with mass losses of (obs. = 9.40%,


calc. = 9.57%). The second phase is from 210 to 450

C and
is attributed to the loss of the organic moiety C
20
H
18
NS with
mass losses of (obs. = 44.37%, calc. = 44.07%). The nal
phase shows the loss of the organic moiety, C
8
H
10
OS, at 450
700

C with mass losses of (obs. = 22.53%, calc. = 22.33%)


leaving Mn as the nal product, and the fragment C
8
N is lost
as 8CO
2
and 0.5N
2
.
The decomposition of the Co(II) complex also occurred
in three steps. The rst step is due to the loss of a
mole of water and CH
4
at 30200

C, with mass losses of


(obs. = 4.77%, calc. = 4.90%). The successive decomposi-
tion occurs within a temperature range of 200400

C and
is attributed to the loss of the organic moiety C
6
H
8
S
2
with
mass losses of (obs. = 21.45%, calc. = 20.76%). The last step
involves the loss of the organic moiety, C
24
H
16
N
2
O
4
, at 400
700

C with mass losses of (obs. = 57.50%, calc. = 57.08%).


The nal product is Co, and the C
5
fragment is lost as 5CO
2
.
The TGA curve of the Ni(II) complex reveals a three-
step decomposition. The rst is the loss of 0.5 mole of water
and SO
2
at 30240

C, with mass losses of (obs. = 10.56%,


calc. = 10.67%). The second step ranges from 240 to 420

C
and is assigned to the loss of the organic moiety, C
10
H
10
N
2
with mass losses (obs. = 22.73%, calc. = 23.08%). The nal
step is within a temperature range of 420700

C and is
attributed to the loss of the organic moiety C
25
H
18
SO
2
(obs. = 55.51%, calc. = 55.81%). The remaining fraction is
Ni residue, and the fragment CH is lost as CO
2
+ 0.5H
2
.
Cu(II) complex decomposes in three steps. The rst step
is attributed to the loss of the fragment C
8
H
12
N, with mass
losses of (obs. = 17.86%, calc. = 17.89%) at 30200

C. The
second step ranges from 200 to 400

C and is attributed
to the loss of the fragment C
13
H
10
O
2
S, with mass losses
of (obs. = 33.63%, calc. = 33.74%). The nal step is from
400 to 700

C corresponding to the loss of the organic


moiety C
15
H
4
O
2
NS, with mass losses of (obs. = 42.0%,
calc. = 38.72%). The remaining residue is Cu.
The Zn(II) complex decomposes in three steps. Step one
is between 30260

C, which indicates the loss of H


2
S, with
mass losses of (obs. = 5.17%, calc. = 5.00%). The second
step involves the loss of the organic moiety C
11
H
10
OS,
6 International Journal of Inorganic Chemistry
from 260 to 440

C, with mass losses of (obs. = 27.80%,


calc. = 27.87%). The nal step is attributed to the loss of the
organic moiety, C
22
H
16
N
2
O
3
, at 440700

C, with mass losses


of (obs. = 52.21%, calc. = 52.22%), leaving behind the Zn
residue, and the C
3
fragment is lost as 3CO
2
.
The Pd(II) complex also decomposes in three phases.
The rst phase is between 30210

C and is attributed
to loss of C
2
H
4
, with mass losses of (obs. = 3.50%,
calc. = 3.87%). The second phase involves the loss of
the organic moiety, C
18
H
15
O
2
S at 210400

C, with mass
losses of (obs. = 41.33%, calc. = 40.79%) while the nal
stage involves the loss of the organic moiety, C
13
H
9
N
2
SO,
from 400 to 700

C, with mass losses of (obs. = 33.42%,


calc. = 33.32%), leaving behind PdO residue, and the frag-
ment C
3
is lost as 3CO
2
.
In all cases with the exceptions of the ligand and the
Cu(II) complex, the decomposition pattern showed the
loss of carbon fragments which got oxidized to CO
2
, and
hydrogen or nitrogen which were lost as gases. Thus, the
decomposition pattern corroborates the proposed formula-
tion of the complex.
3.5. Conductance. The molar conductances of the com-
plexes in nitromethane are below 20.0 ohm
1
cm
2
mol
1
conrming their covalent nature. A value in the range 75
90 ohm
1
cm
2
mol
1
is expected for a 1 : 1 electrolyte [12].
4. Conclusion
The Schi-base ligand coordinates to the Mn(II), Ni(II),
Co(II), Cu(II), Pd(II), and Zn(II) ions in a tetradentate
manner using the N
2
O
2
chromophores. The assignment of
a 4-coordinate square-planar geometry to Cu(II) and Pd(II)
complexes and tetrahedral geometry to Mn(II), Ni(II),
Co(II), and Zn(II) complexes is corroborated by elemental
analysis, thermal, magnetic, and electronic spectral measure-
ments. The Cu(II) and Pd(II) complexes exhibit geometric
isomerism and are in the trans-isomeric form as conrmed
by their infrared spectra. Furthermore, the Cu(II) complex
exhibits some anti-ferromagnetic interactions, operating
through a dimeric structure while the other complexes are
mononuclear.
Acknowledgment
The author thanks The Alexander von Humboldt (AvH)
Foundation for a Georg Forster Fellowship.
References
[1] A. A. Osowole, G. A. Kolawole, R. Kempe, and O. E. Fagade,
Spectroscopic, magnetic and biological studies on some
metal(II) complexes of 3-(4,6-Dimethyl-2Pyrimidinylamino)-
1-Phenyl-2-Butenone and their adducts with 2,2-Bipyridine
and 1,10-Phenanthroline, Synthesis and Reactivity in Inor-
ganic, Metal-Organic and Nano-Metal Chemistry, vol. 39, no.
3, pp. 165174, 2009.
[2] A. A. Osowole, G. A. Kolawole, and O. E. Fagade, Synthesis,
characterization and biological studies on unsymmetrical
Schi-base complexes of nickel(II), copper(II) and zinc(II)
and adducts with 2,2-dipyridine and 1,10-phenanthroline,
Journal of Coordination Chemistry, vol. 61, no. 7, pp. 1046
1055, 2008.
[3] A. A. Osowole, B. C. Ejelonu, and S. A. Balogun, Spec-
troscopic, magnetic and antibacterial properties of some
metal(II) unsymmetric Schi-base complexes and their
mixed-ligand analogs, Journal of Ultra Chemistry, vol. 20, no.
3, pp. 549558, 2008.
[4] A. A. Osowole and O. E. Fagade, Synthesis, characterization
and biopotency of some metal(II) -ketoiminates and their
mixed-ligand complexes, Polish Journal of Chemistry, vol. 81,
no. 12, pp. 20392048, 2007.
[5] A. A. Osowole, G. A. Kolawole, and O. E. Fagade, Synthesis,
physicochemical, and biological properties of nickel(II), cop-
per(II), and zinc(II) complexes of an unsymmetrical tetraden-
tate Schi base and their adducts, Synthesis and Reactivity in
Inorganic, Metal-Organic and Nano-Metal Chemistry, vol. 35,
no. 10, pp. 829836, 2005.
[6] M. H. E. Chan, K. A. Crouse, M. I. M. Tahir, R. Rosli,
N. Umar-Tsafe, and A. R. Cowley, Synthesis and char-
acterization of cobalt(II), nickel(II), copper(II), zinc(II)
and cadmium(II) complexes of benzyl N-[1-(thiophen-
2-yl)ethylidene] hydrazine carbodithioate and benzyl N-
[1-(thiophen-3-yl)ethylidene] hydrazine carbodithioate and
the X-ray crystal structure of bisbenzyl N-[1-(thiophen-
2-yl)ethylidene] hydrazine carbodithioatenickel(II), Polyhe-
dron, vol. 27, no. 4, pp. 11411149, 2008.
[7] P. S. Kenderekar, S. V. More, P. S. Patil, S. RS. Bhusare, and R. P.
Pawar, Synthesis of 2-(2-hydroxy- 3 -iodo-5-bromo phenyl)-
3 -(substituted phenyl)- 4 -thiazolidinones as antibacterial
agents, Oriental Journal of Chemistry, vol. 18, no. 3, pp. 595
597, 2002.
[8] N. Idrees, M. Siddique, A. G. Doshi, and A. W. Raut, Synthe-
sis of N-substituted phenyl-4- thiophenyl -2-azetidinones and
its antimicrobial activity, Oriental Journal of Chemistry, vol.
17, no. 1, pp. 143146, 2001.
[9] N. Idrees, M. Siddique, S. D. Patil, A. G. Doshi, and A. W. Raut,
Synthesis of Schi bases of thiophene-2-carboxaldehyde and
its antimicrobial activity, Oriental Journal of Chemistry, vol.
17, no. 1, pp. 131133, 2001.
[10] B. N. Reddy, P. G. Avaji, P. S. Badami, and S. A. Patil, Synthe-
sis, spectral and biological studies of cobalt(II), nickel(II) and
copper(II) complexes with 1,5-bis(thiophenylidene) thiocar-
bohydrazone, Journal of Saudi Chemical Society, vol. 11, no.
2, pp. 253268, 2007.
[11] Z. H. Chohan, H. Pervez, K. M. Khan, A. Rauf, and C. T. Supu-
ran, Binding of transition metal ions [cobalt, copper, nickel
and zinc] with furanyl-, thiophenyl-, pyrrolyl-, salicylyl- and
pyridyl-derived cephalexins as potent antibacterial agents,
Journal of Enzyme Inhibition and Medicinal Chemistry, vol. 19,
no. 1, pp. 5156, 2004.
[12] Z. H. Chohan, H. Pervez, K. M. Khan, A. Rauf, G. M.
Maharvi, and C. T. Supuran, Antifungal cobalt(II), cop-
per(II), nickel(II) and zinc(II) complexes of furanyl-thio-
phenyl-, pyrrolyl-, salicylyl- and pyridyl-derived cephalexins,
Journal of Enzyme Inhibition and Medicinal Chemistry, vol. 19,
no. 1, pp. 8590, 2004.
[13] Z. H. Chohan, H. Pervez, A. Rauf, A. Scozzafava, and
C. T. Supuran, Antibacterial Co(II), Cu(II), Ni(II) and
Zn(II) complexes of thiadiazole derived furanyl, thiophenyl
and pyrrolyl schi bases, Journal of Enzyme Inhibition and
Medicinal Chemistry, vol. 17, no. 2, pp. 117122, 2002.
International Journal of Inorganic Chemistry 7
[14] Z. H. Chohan and S. Kausar, Synthesis, characterization
and biological properties of tridentate NNO, NNS and NNN
donor thiazole-derived furanyl, thiophenyl and pyrrolyl Schi
bases and their Co(II), Cu(II), Ni(II) and Zn(II) metal
chelates, Metal-Based Drugs, vol. 7, no. 1, pp. 1722, 2000.
[15] A. I. P. Sinha and M. Bala, Coordination behavior of herbici-
dal Schi bases derived from 2-amino-6-ethoxybenzothiazole
towards copper(II), Asian Journal of Chemistry, vol. 3, no. 1,
pp. 4551, 1991.
[16] P. Mittal, S. Joshi, V. Panwar, U. Vatsala, and V. Uma,
Biologically active Co (II), Ni (II), Cu (II) and Mn(II)
Complexes of Schi bases derived from vinyl aniline and
heterocyclic aldehydes, International Journal of ChemTech
Research, vol. 1, no. 2, pp. 225232, 2009.
[17] M. Ghazzali, V. Langer, C. Lopes, A. Eriksson, and L.

Ohrstr om, Syntheses, crystal structures, optical limiting


properties, and DFT calculations of three thiophene-2-
aldazine Schi base derivatives, NewJournal of Chemistry, vol.
31, no. 10, pp. 17771784, 2007.
[18] D. L. Wang, W. S. Hwang, L. C. Liang, L. I. Wang, L. Lee,
and M. Y. Chiang, Reaction of a thienyl schi base with
diiron nonacarbonyl: characterization and structures of [-N-
(((2,3-: )-5-methyl-2-thienyl)methyl)-: (N)-anilino]hex-
acarbonyldiiron and [-N-((anilino(2-thienyl)methyl) ((2,3-
:)-2-thienyl)methyl)- :(N), Organometallics, vol. 16, no.
14, pp. 31093113, 1997.
[19] J. Wiedermann, K. Mereiter, and K. Kirchner, Palladium
imine and amine complexes derived from 2-thiophenecar-
boxaldehyde as catalysts for the Suzuki cross-coupling of aryl
bromides, Journal of Molecular Catalysis A, vol. 257, no. 1-2,
pp. 6772, 2006.
[20] J. Bassett, R. C. Denney, G. H. Jeery, and J. Mendham, Vogels
Textbook of Quantitative Inorganic Analysis, ELBS, London,
UK, 1978.
[21] A. Earnshaw, The Introduction to Magnetochemistry, Academic
Press, London, UK, 1980.
[22] A. A. Nejo, G. A. Kolawole, A. R. Opoku, C. Muller, and
J. Wolowska, Synthesis, characterization, and insulin-en-
hancing studies of unsymmetrical tetradentate Schi-base
complexes of oxovanadium(IV), Journal of Coordination
Chemistry, vol. 62, no. 21, pp. 34113424, 2009.
[23] A. B. Lever, Inorganic Electronic Spectroscopy, Elsevier, London,
UK, 1980.
[24] E. Kwiatkowski and M. Kwiatkowski, Unsymmetrical schi
base complexes of nickel(II) and palladium(II), Inorganica
Chimica Acta, vol. 42, no. C, pp. 197202, 1980.
[25] M. G. B. Drew and P. C. Yates, Synthetic, structural and
molecular mechanics investigation of some mono- and din-
uclear copper(I) and copper(II) complexes of macrocyclic and
related acyclic ligands, Inorganica Chimica Acta, vol. 118, no.
1, pp. 3747, 1986.
[26] S. Cherayath, J. Alice, and C. P. Prabhakaran, Palladi-
um(II) complexes of Schi bases derived from 5-amino-2,4-
(1H, 3H)pyrimidinedione (5-aminouracil) and 1,2-dihydro-
1,5-dimethyl-2-phenyl-4-amino-3H-pyrazol-3-one, Transi-
tion Metal Chemistry, vol. 15, no. 6, pp. 449453, 1990.
[27] X. Han, Z. L. You, Y. T. Xu, and X. M. Wang, Synthesis, char-
acterization and crystal structure of a mononuclear zinc(II)
complex derived from 2-methoxy- 6-[(3-cyclohexylamino-
propylimino)methyl]phenol, Journal of Chemical Crystallog-
raphy, vol. 36, no. 11, pp. 743746, 2006.
Hindawi Publishing Corporation
International Journal of Inorganic Chemistry
Volume 2011, Article ID 606271, 9 pages
doi:10.1155/2011/606271
Research Article
Synthesis and Structural Characterization of
a NewTetranuclear Nickel(II) Sulfato Complex Containing
the Anionic Formof Di-2-Pyridyl Ketone Oxime
Eleni Moushi,
1
Constantinos G. Efthymiou,
2
Spyros P. Perlepes,
3
and Constantina Papatriantafyllopoulou
2, 3
1
Department of Chemistry, University of Cyprus, 1678 Nicosia, Cyprus
2
Department of Chemistry, University of Florida, Gainesville, FL 32611-7200, USA
3
Department of Chemistry, University of Patras, 265 04 Patras, Greece
Correspondence should be addressed to Constantina Papatriantafyllopoulou, cpapat@chem.u.edu
Received 7 December 2010; Accepted 26 January 2011
Academic Editor: Daniel L. Reger
Copyright 2011 Eleni Moushi et al. This is an open access article distributed under the Creative Commons Attribution License,
which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.
The preparation and crystal structure of a tetranuclear Ni(II) sulfato cluster containing the anion of di-2-pyridyl ketone oxime,
(py)
2
CNO

, are reported. Treatment of NiSO


4
6H
2
O with one equivalent of (py)
2
CNOH and one equivalent of NEt
3
in MeOH
leads to the compound [Ni
4
{(py)
2
CNO}
4
(SO
4
)
2
(MeOH)
4
] (1) in moderate yield. The metal ions are linked together by two 3.2111
and two 2.1110 (Harris notation) (py)
2
CNO

ligands, as well as two 2.1100 SO


4
2
ions to create a rare metallacrown-type (12-
MC-4) ring. Strong H-bond intermolecular interactions in 1 lead to the formation of a 1D chain along the c axis. Characteristic
IR bands are discussed in terms of the known structure of 1.
1. Introduction
There is currently a renewed interest in the coordination
chemistry of oximes [120]. 2-pyridyl oximes (Scheme 1)
are popular ligands in coordination chemistry [2136]. The
anions of these molecules are versatile ligands for a variety
of research objectives, including
2
and
3
behaviour [21,
22]; the activation of 2-pyridyl oximes by 3d-metal centers
towards further reactions is also becoming a fruitful area of
research [21, 22, 26]. The majority of the metal complexes
of these ligands have been prepared in the last 15 years and
much of their chemistry remains to be explored in more
detail [22].
We have been exploring ligand blend reactions involv-
ing carboxylates (R

CO
2

) and various 2-pyridyloximates


with (ternary ligand blends) or without (binary ligand
blends) additional inorganic monoanions (Cl

, Br

, NO
3

,
N
3

, SCN

) as a means to high-nuclearity species. The


presence of a deprotonated oxime group leads to a great
coordinative exibility due to the well-known ability of the
oximate group to bridge two or three metal ions. On the
other hand, carboxylates are able to deprotonate the oxime
group of 2-pyridyloximes under mild conditions (the use of
external hydroxides often perplexes the reactions). Besides
their deprotonating ability, the R

CO
2

ions are exible


ligands, a consequence of their ability to adopt a number of
dierent ligation modes, both terminal and bridging as well
as both bidentate and tridentate. The additional inorganic
monoanions in the ternary ligand blends often behave as
terminal ligands and help the formation of clusters (and
not coordination polymers). However, sometimes they act
as bridging ligands, and this may eventually lead to clusters
with complicated structures; the formation of coordination
polymers cannot be ruled out in such a case. Thus, a
variety of Cr, Mn, Fe, Co, Ni, and Cu clusters [2239] with
nuclearities ranging from 3 to 12 have been characterized
from our [2336] and other [21, 3739] groups, some of
them possessing interesting magnetic properties, including
single-molecule magnetism behaviour [33, 40].
Recently, we have begun a program which can be
considered as a modication of the above-mentioned binary
ligand blend approach. We have been exploring the
use of other inorganic ions, such as SO
4
2
, instead of
the carboxylato ligand, R

CO
2

, in the 3d-metal cluster


2 International Journal of Inorganic Chemistry
N
C
R
N
OH
R = H; (py)C(H)NOH
R = Me; (py)C(Me)NOH
R = Ph; (py)C(Ph)NOH
R =
N
; (py)
2
CNOH
Scheme 1: General structural formula and abbreviations of simple
2-pyridyl oximes.
chemistry with 2-pyridyloximate ligands. The sulfate ion
[41] is a ligand with great coordinative exibility (
2
,
3
,

4
,
5
,
6
,
8
, or
10
potential), see Scheme 2. Metal-sulfato
complexes have been studied for their roles in the eld of
porous framework materials [42, 43], in catalysis [44], in the
construction of luminescent molecular materials [45, 46],
and in medicinal [47], environmental [48], and bioinorganic
[49] chemistry. The possible advantages of using SO
4
2
instead of R

CO
2

include (i) the possibility of triggering


aggregation of preformed smaller cationic species into new,
higher-nuclearity products and (ii) the possible diversion
of known reaction systems developed using monoanionic
carboxylates to new species as a result of the higher charge
and higher denticity/bridging capability of sulfates. Thus, the
initial employment of the sulfate ion in Ni
II
/(py)C(R)NOH
(R = Me, Ph, NH
2
) chemistry has led to the isolation and
characterization of high-nuclearity Ni
II
compounds, such as
Ni
12
[50] and Ni
6
[51, 52] clusters which possess interesting
structural properties.
In this work, we expand our eorts to a dier-
ent member of 2-pyridyl oximes which is di-2-pyridyl
ketone oxime, (py)
2
CNOH, and report the synthesis
and characterization of the new tetranuclear compound
[Ni
4
{(py)
2
CNO}
4
(SO
4
)
2
(MeOH)
4
]. The structure of the
compound has been determined by single-crystal X-ray
diraction. The IR data are discussed in terms of the nature
of bonding and the structure of the complex.
2. Experimental
2.1. General and Physical Measurements. All manipulations
were performed under aerobic conditions using materials
(reagent grade) and solvents as received.
Microanalyses (C, H, N) were performed by the Univer-
sity of Ioannina (Greece) Microanalytical Laboratory using
an EA 1108 Carlo Erba analyzer. IR spectra (4000-400 cm
1
)
were recorded on a Perkin-Elmer 16 PC FT-spectrometer
with samples prepared as KBr pellets.
2.2. Compound Preparation
2.2.1. [Ni
4
{(py)
2
CNO}
4
(SO
4
)
2
(MeOH)
4
] (1). NEt
3
(0.139
ml, 1.00 mmol) was added to a colourless solution of
(py)
2
CNOH (0.199 g, 1.00 mmol) in MeOH (25 ml). Subse-
quently, solid NiSO
4
6H
2
O (0.263 g, 1.00 mmol) was added,
and the resulting red solution was stirred for 1 h at room
temperature. A small quantity of undissolved material was
Table 1: Summary of crystal data, data collection, and structure
renement for the X-ray diraction study of complex 1.
Complex 1
Empirical formula Ni
4
C
48
H
48
N
24
O
16
S
2
Formula weight 1515.97
Colour, habit orange, rod
Crystal system orthorhombic
Space group P
ccn
a (

A) 14.2969 (10)
b (

A) 20.9678 (14)
c (

A) 18.5395 (13)
V (

A
3
) 5557.7 (7)
Z 4

calc
(g cm
3
) 1.609
Radiation, (

A) 0.71073
(mm
1
) 1.488
F (000) 2760
Temperature (K) 100 (2)
2
max
[

] 61
Ranges h 18 18
k 21 27
l 27 23
Measured reections 42848
Unique reections 6377
Reections used 3703
(I > 2(I))
Parameters rened 374
GoF (on F
2
) 0.903
R
1
a
(I > 2(I)) 0.0464
w R
2
b
(I > 2(I)) 0.1042
()
max
/()
min
(e

A
3
) 0.861/0.621
a
R
1
=

(|F
o
| |F
c
|)/

(|F
o
|).
b
wR
2
= {

[w(F
o
2
F
c
2
)
2
]/

[w(F
o
2
)
2
]}
1/2
.
removed by ltration and the dark red ltrate layered
with Et
2
O (50 ml). Slow mixing gave X-ray quality, orange
crystals which were collected by ltration, washed with
Et
2
O (2 3 ml), and dried in air; yield 57%. The dried
solid was analyzed satisfactorily as 1MeOH. Anal. Calc. for
C
49
H
52
Ni
4
N
24
O
17
S
2
: C, 38.02; H, 3.99; N, 21.72. Found: C,
38.45; H, 3.87; N, 21.37%. IR (KBr pellet): ` v = 3388 sb,
2902 w, 1654 w, 1598 m, 1460 m, 1430 m, 1376 w, 1340 w,
1282 w, 1219 m, 1130 m, 1118 s, 1086 s, 1045 s, 1020 m,
982 m, 896 w, 788 w, 748 m, 702 m, 670 m, 641 m, 618 s,
591 w, 452 wcm
1
.
2.3. Single-Crystal X-Ray Crystallography. A crystal of 1 with
appropriate dimensions 0.08 0.03 0.01 mm was attached
to a glass ber using silicone grease. Data were collected on
an Oxford Diraction Xcalibur-3 diractometer, equipped
with a Sapphire CCD area detector, at 100 K, using a graphite
monochromated Mo K radiation. Complete crystal data
and parameters for data collection and processing are listed
in Table 1.
International Journal of Inorganic Chemistry 3
S
O
O
O
O
M
M M
M
M
M
M
M
S
O
O
O
O
M
M
M
M
M
M
M
M
M
M
8.2222
5
3.1110
S
O
O
O
O
M
S
O
O
O
M
O
S
O
O
O
O
M
M
S
O
O
O
O
M
M
S
O
O
O
O
M
M
M
M
O
S
O
O
O
M
M
M
M
M
O
S O
O
O
M M
O
S
O
O
O
M
M
M
M
M
M
1.1000 2.2000
4.1111
5.2210
2.2110
6.2211
O
S
O
O
O
M
M
M
3.2100
O
S
O
O
O
M
M
M
3.2111
O
S
O
O
O
M
M
M
M
M
5.2211
S
O
O
O
O
M
M
M
M
4.2210
O
S
O
O
O
M
M
M
S
O
O
O
O
M
M
2.1110
O
S
O
O
O
M
M
M
M
4.2200
S
O
O
O
M
O
M
2.1111
2.1100 1.1100
10.3322
Scheme 2: The up to now crystallographically established coordination modes of the sulfato ligand and the Harris notation [53] which
describes these modes.
The structure was solved by direct methods using SIR92
[54] and rened by full-matrix least-squares techniques on
F
2
with SHELXL-97 [55]. Some residual electron density
in the accessible voids of the structure was too disordered
to rene as solvent molecules; therefore, the SQUEEZE
procedure [56] of PLATON was employed to remove the
contribution of the electron density in the solvent region
from the intensity data. The solvent-free model and intensity
data were used for the nal results reported here. The non-
H atoms were treated anisotropically. The H atoms of the
(py)
2
CNOH ligands and the methyl groups of the methanol
molecules were placed in calculated, ideal positions and
4 International Journal of Inorganic Chemistry
N4
N5
N6
N3
O1
O2
Ni2
O3
O4
O7
N2
N1
O8
Ni1
N6

O2

N5

O8

O1

Ni1

O4

O3

N3

N2

Ni2

O7

N4

N1

Figure 1: The molecular structure of 1. H atoms have been omitted


for clarity.
rened as riding on their respective C atoms. The H atom
of the OH group of one independent methanol molecule
(O(8)H) was located in dierence Fourier maps and was
rened isotropically, but the H atom of the OH group of the
second independent methanol molecule (O(7)H) could not
be located. The programs used were CRYSALIS CCD [57]
for data collection, CRYSALIS RED [57] for cell and data
renement, WINGX [58] for crystallographic calculations,
and MERCURY [59] and DIAMOND [60] for molecular
graphics.
3. Results and Discussion
3.1. Synthetic Comments. Our general synthetic approach
for the isolation of Ni
II
/2-pyridyloximate/sulfato clusters has
been to treat the metal sulfate salt with the appropriate
ligand and a base in a variety of solvents. The addition of
base is necessary for the deprotonation of the oxime ligand.
Treatment of NiSO
4
6H
2
O with one equivalent of
(py)
2
CNOH and one equivalent of NEt
3
in MeOH gave a
red solution which, upon crystallization, gave orange crystals
of the new tetranuclear cluster which can be written as
[Ni
4
{(py)
2
CNO}
4
(SO
4
)
2
(MeOH)
4
] (1). Its formation can be
summarized in (1)
4NiSO
4
6H
2
O + 4
_
py
_
2
CNOH + 4NEt
3
+ 4MeOH
MeOH

_
Ni
4
_
_
py
_
2
CNO
_
4
(SO
4
)
2
(MeOH)
4
_
+ 2(HNEt
3
)
2
SO
4
+ 24H
2
O.
(1)
As expected, the nature of the base is not crucial for the
identity of the product, and it aects only its crystallinity,
and in some cases its purity; we were able to isolate 1 by
using a plethora of dierent bases such as NaOMe, NMe
4
OH,
N5
O2
Figure 2: Top: the [Ni
4
(-SO
4
)
2
(
2
-ONR)
2
(
3
-ONR)
2
] core of
complex 1, where R- = -C(py)
2
. The metallacrown-type ring is
highlighted. Bottom: the core of 1 using only the
3
oximate groups.
NEt
4
OH, and LiOHH
2
O. Small changes in the molar ratio
of the reactants, the crystallization method, and the presence
of counterions do not seem to aect the identity of the
isolated product.
3.2. Description of Structure. Partially labeled plots of
the complete structure and the core of the molecule
[Ni
4
{(py)
2
CNO}
4
(SO
4
)
2
(MeOH)
4
] that is present in com-
plex 1 are shown in Figures 1 and 2, respectively. Selected
interatomic distances and angles are listed in Table 2.
The structure of 1 consists of tetranuclear molecules
[Ni
4
{(py)
2
CNO}
4
(SO
4
)
2
(MeOH)
4
] which lie on a crystal-
lographic inversion center. The metal ions are held together
by two 3.2111 and two 2.1110 (using Harris notation, [53],
Scheme 3) (py)
2
CNO

ligands, as well as two 2.1100 SO


4
2
ions. Four MeOH molecules act as terminal ligands and
complete the coordination sphere of the four metal centers.
The molecule has a metallacrown-type topology [61]. A
pseudo 12-MC-4 ring forms; the true 12-MC-4 topology is
destroyed by the bridging character of the oximate oxygen
atoms O2 and O2

.
Adistorted octahedral environment is created about each
metal center; the chromophores are represented by the fol-
lowing formulas Ni(1,1

)(N
py
)(N
ox
)(O
ox
)
2
(O
sulf
)(O
met
) and
Ni(2,2

)(N
py
)
2
(N
ox
)(O
ox
)(O
sulf
)(O
met
), where the abbrevia-
tions py, ox, sulf, and met are for the 2-pyridyl,
oximate, sulfato, and methanolic donor atoms, respectively.
The average NiO
ox
, NiN
ox
, and NiN
py
bond lengths
of 2.054(3), 2.043(3), and 2.071(3)

A, respectively, agree
well with the values expected for high-spin Ni
II
ions in
octahedral environment [34, 35, 6264]. The NiO
sulf
bond
lengths are typical [41, 51, 62, 65]. The fact that S-(O3, O4)
(average 1.474

A) > S-(O5, O6) (average 1.450

A) reects the
coordinating nature of O3 and O4 and the noncoordinating
International Journal of Inorganic Chemistry 5
Table 2: Selected interatomic distances (

A) and angles () for 1.


Ni1-O1 2.046 (3) Ni2-O2 2.071 (3)
Ni1-O2 2.046 (3) Ni2-O3 2.042 (3)
Ni1-O4 2.033 (3) Ni2-O7 2.163 (3)
Ni1-O8 2.048 (3) Ni2-N1 2.090 (3)
Ni1-N5 2.090 (3) Ni2-N2 1.997 (4)
Ni1-N6 2.074 (3) Ni2-N4 2.053 (3)
O1-Ni1-O2 89.59 (11) O2-Ni2-O3 94.82 (11)
O1-Ni1-O4 101.65 (11) O2-Ni2-O7 88.69 (10)
O1-Ni1-O8 173.53 (11) O2-Ni2-N1 169.13 (12)
O1-Ni1-N5 82.89 (11) O2-Ni2-N2 90.44 (12)
O1-Ni1-N6 83.38 (12) O2-Ni2-N4 84.07 (12)
O2-Ni1-O4 84.24 (11) O3-Ni2-O7 176.29 (10)
O2-Ni1-O8 88.62 (11) O3-Ni2-N1 88.91 (12)
O2-Ni1-N5 107.12 (11) O3-Ni2-N2 95.11 (12)
O2-Ni1-N6 170.67 (12) O3-Ni2-N4 87.08 (12)
O4-Ni1-O8 84.36 (12) O7-Ni2-N1 87.86 (11)
O4-Ni1-N5 167.92 (13) O7-Ni2-N2 86.05 (12)
O4-Ni1-N6 91.16 (12) O7-Ni2-N4 92.09 (12)
O8-Ni1-N5 91.70 (12) N1-Ni2-N2 79.04 (13)
O8-Ni1-N6 99.02 (12) N1-Ni2-N4 106.36 (13)
N5-Ni1-N6 78.16 (13) N2-Ni2-N4 174.25 (13)
N
C
N
N
O Ni
Ni
Ni
N
C
N
N
O
Ni
Ni
3.2111 2.1110
Scheme 3: The two dierent coordination modes of the (py)
2
CNO

ligands present in complex 1 and the Harris notation [53] which


describes them.
character of O5 and O6; as expected, the sulfur to free
oxygen bond lengths are the shortest.
The crystal structure of 1 is stabilized by strong inter-
and intramolecular hydrogen bonds. The intramolecular
hydrogen bonds involve the O atom (O7 and its symmetry
equivalent) belonging to a methanol ligand as donor and
the O atom (O1 and its symmetry equivalent) of the
doubly bridging organic ligand as acceptor [O1 O7 =
2.721(3)

A]. The O atom of the remaining methanol ligand
(O8 and its symmetry equivalent) is participating as donor
in an intermolecular hydrogen bond with the acceptor being
the pyridyl N atom (N3 and its symmetry equivalent) of the
doubly bridging organic ligand [O8 N3 = 2.851(3)

A,
H(O8) N3 = 1.987(3)

A, and O8-H(O8)-N3 = 172.1(1)].
This hydrogen bonding leads to the formation of a 1D chain
along the c axis (Figure 3).
Figure 3: Representation of a part of the 1Dchain formed in 1 along
the c axis. The dotted lines represent H bonds.
The molecule of 1 contains the [Ni
4
(-SO
4
)
2
(
2
-
ONR)
2
(
3
-ONR)
2
] core, where R- =-C(py)
2
(Figure 2, top).
An alternative description of the core (using only the
3
oximate groups) is [Ni
4
(
3
-ONR)
2
]
6+
(Figure 2, bottom).
The topology of the four Ni
II
ions can be also described
as saddle-like, and it is observed for the rst time in Ni
4
clusters. The most common topologies of the metal ions in
Ni
II
4
complexes are the cubanes [6673] and the face-shared
distorted dicubanes in which one of the corners of each
cubane is missing [7479], while there are few Ni
4
clusters
in which the metal ions adopt less common topologies such
as linear [8082], rectangular [8387], and chair-like [88, 89]
as follows.
Complex 1 joins a small but growing family of struc-
turally characterized Ni(II) complexes containing the neutral
or anionic forms of di-2-pyridyl ketone oxime as ligands
[34, 36, 8892]. The special features of 1 compared to the
other members of this family are (1) It is the rst example
of these species containing the sulfato ligand, and (2) it has a
unique Ni
4
clusters saddle-like metal topology.
3.3. IR Spectra. The medium intensity bands at 1568 and
1094 cm
1
in the spectrum of the free ligand (py)
2
CNOH are
assigned to v(C=N)
oxime
and v(N-O)
oxime
modes, respectively
[51, 52, 93]. The 1094 cm
1
band is shifted to a higher
wavenumber (1118 cm
1
) in 1. This shift is in accord
with the concept that upon deprotonation and oximate-
O coordination, there is a higher contribution of N=O to
the electronic structure of the oximate group; consequently,
the v(N-O) vibration shifts to a higher wavenumber in
the complex relative to (py)
2
CNOH [36]. Somewhat to
our surprise, the 1568 cm
1
band is shifted to a higher
wavenumber in the complex (1598 cm
1
), overlapping with
an aromatic stretch. This shift may be indicative of the oxime
nitrogen coordination [94]. Extensive studies on Schi base
complexes (which also contain a C=N bond) have shown
[95] that a change in the s character of the nitrogen lone
pair occurs upon coordination such that the s character of
nitrogen orbital involved in the C=N bond increases; this
change in hybridization produces a greater C=N stretching
force constant relative to the free neutral ligand.
The in-plane deformation band of the 2-pyridyl ring
of free (py)
2
CNOH at 622 cm
1
shifts upwards (641 cm
1
),
conrming the involvement of the ring-N atom in coordina-
tion [96]. The presence of the 618 cm
1
bond in the spectrum
of 1 indicates that some 2-pyridyl rings are free, that is,
6 International Journal of Inorganic Chemistry
uncoordinated, in accordance with the 2.1110 (py)
2
CNO

ligands that are present in the complex.


The IR spectrum of the free, that is, ionic, sulfate (the
ion belongs to the T
d
point group) consists of two bands
at 1105 and 615 cm
1
, assigned to the v
3
(F
2
) stretching
[
d
(SO)] and v
4
(F
2
) bending [
d
(OSO)] modes, respectively
[41, 97]. The v
1
(A
1
) stretching [
s
(SO)] and v
2
(E) bending
[
d
(OSO)] fundamentals are not IR active. The coordination
of SO
4
2
to metal ions decreases the symmetry of the group,
and the v
3
and v
4
modes are split [41, 97]. In the case
when the SO
4
2
-site symmetry is lowered from T
d
to C
2v
(bidentate chelating or bridging coordination), which is the
case in 1, both v
1
and v
2
appear in the IR spectrum, while
v
3
and v
4
each splits into three IR-active vibrations [97].
Thus, the bands at 1219, 1130, and 1020 cm
1
are attributed
to the v
3
modes [97], while the bands at 591, 618, and
670 cm
1
are assigned to the v
4
modes [13, 7479] with
the intermediate wavenumber band being superimposed by
a ligands vibration. The band at 982 cm
1
and the weak
feature at 452 cm
1
can be assigned [41, 97] to the v
1
and
v
2
modes, respectively. These spectral features agree with the
low C
2v
symmetry for the sulfato ligand in the complex, as
also conrmed crystallographically.
4. Conclusions
The present work extends the body of results that emphasize
the ability of the sulfate ion to create unique structural
types in 3d-metal cluster chemistry. The study of the
coordination chemistry of the binary SO
4
2
/(py)
2
CNOH
ligand system in the presence of base in MeOH has
provided access to the novel tetranuclear Ni(II) cluster
[Ni
4
{(py)
2
CNO}
4
(SO
4
)
2
(MeOH)
4
] (1). Complex 1 contains
the [Ni
4
(-SO
4
)
2
(
2
-ONR)
2
(
3
-ONR)
2
] core, where R- = -
C(py)
2
, with a unique saddle-like topology of the Ni
II
ions; it
is thus a valuable addition to the family of tetranuclear Ni
II
clusters.
Analogues of 1 with2-pyridinealdoxime [(py)C(H)NOH],
methyl(2-pyridyl)ketone oxime [(py)C(Me)NOH], or
phenyl(2-pyridyl)ketone oxime [(py)C(ph)NOH] (Scheme
1) are not known, until to date, and further research eorts
are in progress to determine the appropriate reaction
conditions that could possibly favor such species. It is
likely that the preparation and stability of such tetranuclear
complexes are dependent on the particular nature of the R
substituent on the oximate carbon. We are currently working
on the chemistry of the NiSO
4
6H
2
O/(py)C(R)NOH (R=H,
Me, Ph) reaction systems.
Supporting Information
CCDC 802606 contains the supplementary crystallographic
data for 1. These data can be obtained free of charge via
http://www.ccdc.cam.ac.uk/conts/retrieving.html or from
the Cambridge Crystallographic Data Centre, 12 Union
Road, Cambridge CB2 1EZ, UK; fax: (+44)1223-336033 or
e-mail: deposit@ccdc.cam.ac.uk.
References
[1] V. Yu. Kukushkin and A. J. L. Pombeiro, Oxime and oximate
metal complexes: unconventional synthesis and reactivity,
Coordination Chemistry Reviews, vol. 181, no. 1, pp. 147175,
1999.
[2] A. G. Smith, P. A. Tasker, and D. J. White, The structures
of phenolic oximes and their complexes, Coordination Chem-
istry Reviews, vol. 241, no. 1-2, pp. 6185, 2003.
[3] M. G. Davidson, A. L. Johnson, M. D. Jones, M. D. Lunn,
and M. F. Mahon, Titanium(IV) complexes of oximesnovel
binding modes, Polyhedron, vol. 26, no. 5, pp. 975980, 2007.
[4] D. Mandal, V. Bertolasi, G. Arom, and D. Ray, A ketone
oximate based cyclic cationic [Ni] inverse metallacrown from
simultaneous chelation and bridging of two ligands, Dalton
Transactions, no. 20, pp. 19891992, 2007.
[5] C. P. Raptopoulou, A. K. Boudalis, Y. Sanakis et al., Hex-
anuclear iron(III) salicylaldoximato complexes presenting
the [Fe
6
(
3
-O)
2
(
2
-O)
2
]
12+
core: syntheses, crystal structures,
and spectroscopic and magnetic characterization, Inorganic
Chemistry, vol. 45, no. 5, pp. 23172326, 2006.
[6] V. V. Pavlishchuk, S. V. Kolotilov, A. W. Addison et al., Struc-
tural, magnetic and related attributes of some oximatebridged
tetranuclear nickel(II) rhombs and a dinuclear congener,
Dalton Transactions, no. 8, pp. 15871595, 2003.
[7] D. T. Rosa, J. A. Krause Baue, and M. J. Baldwin, Structural
and spectroscopic studies of the versatile coordination chem-
istry of the chiral ligand N,N-bis(1-propan-2-onyl oxime)-
L-methionine N-methylamide with Ni and Zn, Inorganic
Chemistry, vol. 40, no. 7, pp. 16061613, 2001.
[8] M. J. Goldcamp, S. E. Robison, J. A. Krause Bauer, and M.
J. Baldwin, Oxygen reactivity of a Nickel(II)-polyoximate
complex, Inorganic Chemistry, vol. 41, no. 9, pp. 23072309,
2002.
[9] M. N. Kopylovich, V. Yu. Kukushkin, M. Haukka, J. J. R.
Fra usto da Silva, and A. J. L. Pombeiro, Zinc(II)/ketoxime
system as a simple and ecient catalyst for hydrolysis of
organonitriles, Inorganic Chemistry, vol. 41, no. 18, pp. 4798
4804, 2002.
[10] S. Akine, T. Taniguchi, T. Saiki, and T. Nabeshima, Ca- and
Ba-selective receptors based on site-selective transmetalation
of multinuclear polyoxime-zinc(II) complexes, Journal of the
American Chemical Society, vol. 127, no. 2, pp. 540541, 2005.
[11] A. J. L. Pombeiro and V. Yu. Kukushkin, Reactions of coor-
dinated nitriles, in Comprehensive Coordination Chemistry II,
J. A. McClevery and T. C. Meyer, Eds., vol. 1, pp. 631637,
Elsevier, Amsterdam, The Netherlands, 2004.
[12] J. M. Thorpe, R. L. Beddoes, D. Collison et al., Surface
coordination chemistry: corrosion inhibition by tetranuclear
cluster formation of iron with salicylaldoxime, Angewandte
ChemieInternational Edition, vol. 38, no. 8, pp. 11191121,
1999.
[13] J. Szymanowski, Hydroxyoximes and Copper Hydrometallurgy,
CRC, London, UK, 1993.
[14] I. O. Fritsky, J. Swiatek-Kozlowska, A. Dobosz, T. Yu. Sliva, and
N. M. Dudarenko, Hydrogen bonded supramolecular struc-
tures of cationic and anionic module assemblies containing
square-planar oximate complex anions, Inorganica Chimica
Acta, vol. 357, no. 12, pp. 37463752, 2004.
[15] S. Khanra, T. Weyherm uller, E. Rentschler, and P. Chaudhuri,
Self-assembly of a nonanuclear nickel(II) complex and its
magnetic properties, Inorganic Chemistry, vol. 44, no. 23, pp.
81768178, 2005.
International Journal of Inorganic Chemistry 7
[16] C. J. Milios, C. P. Raptopoulou, A. Terzis et al., Hexanu-
clear manganese(III) single-molecule magnets, Angewandte
ChemieInternational Edition, vol. 43, no. 2, pp. 210212,
2003.
[17] J. P. Costes, F. Dahan, and A. Dupuis, Is ferromagnetism
an intrinsic property of the Cu
II
/Gd
III
couple? 2. Structures
and magnetic properties of novel trinuclear complexes with -
phenolato--oximato (Cu-Ln-Cu) cores (Ln = La, Ce, Gd),
Inorganic Chemistry, vol. 39, no. 26, pp. 59946000, 2000.
[18] D. Robertson, J. F. Cannon, and N. Gerasimchuk, Double-
stranded metal-organic networks for one-dimensional mixed
valence coordination polymers, Inorganic Chemistry, vol. 44,
no. 23, pp. 83268342, 2005.
[19] C. J. Milios, A. Vinslava, W. Wernsdorfer et al., Spin switching
via targeted structural distortion, Journal of the American
Chemical Society, vol. 129, no. 20, pp. 65476561, 2007.
[20] H. Miyasaka, R. Cl erac, K. Mizushima et al.,
[Mn
2
(saltmen)
2
Ni(pao)
2
(L)
2
](A)
2
with L = Pyridine, 4-
picoline, 4-tert-butylpyridine, N-methylimidazole and A=
ClO
4
-, BF
4
-, PF
6
-, ReO
4
-: a family of single-chain magnets,
Inorganic Chemistry, vol. 42, no. 25, pp. 82038213, 2003.
[21] P. Chaudhuri, Homo- and hetero-polymetallic exchange
coupled metal-oximates, Coordination Chemistry Reviews,
vol. 243, no. 1-2, pp. 143190, 2003.
[22] C. J. Milios, T. C. Stamatatos, and S. P. Perlepes, The
coordination chemistry of pyridyl oximes, Polyhedron, vol.
25, no. 1, pp. 134194, 2006.
[23] C. J. Milios, E. Kefalloniti, C. P. Raptopoulou et al.,
Octanuclearity and tetradecanuclearity in manganese
chemistry: an octanuclear manganese(II)/(III) complex
featuring the novel [Mn
8
(
4
-O)
2
(
3
-OH)
2
]
14+
core and
[Mn
II
10
Mn
III
4
O
4
(O
2
CMe)
20
{(2 py)
2
C(OH)O}
4
](2py = 2
pyridyl), Chemical Communications, vol. 9, no. 7, pp. 819
821, 2003.
[24] C. J. Milios, C. P. Raptopoulou, A. Terzis, R. Vicente, A. Escuer,
and S. P. Perlepes, Di-2-pyridyl ketone oxime in 3d-metal
carboxylate cluster chemistry: a new family of mixed-valence
MnMn complexes, Inorganic Chemistry Communications, vol.
6, no. 8, pp. 10561060, 2003.
[25] C. J. Milios, T. C. Stamatatos, P. Kyritsis et al., Phenyl
2-pyridyl ketone and its oxime in manganese carboxylate
chemistry: synthesis, characterisation, X-ray studies and mag-
netic properties of mononuclear, trinuclear and octanuclear
complexes, European Journal of Inorganic Chemistry, no. 14,
pp. 28852901, 2004.
[26] C. J. Milios, E. Kefalloniti, C. P. Raptopoulou et al., 2-
Pyridinealdoxime [(py)CHNOH] in manganese(II) carboxy-
late chemistry: mononuclear, dinuclear, tetranuclear and poly-
meric complexes, and partial transformation of (py)CHNOH
to picolinate(-1), Polyhedron, vol. 23, no. 1, pp. 8395, 2004.
[27] M. Alexiou, E. Katsoulakou, C. Dendrinou-Samara et al.,
Di-2-pyridyl ketone oxime in zinc chemistry: inverse 12-
metallacrown-4 complexes and cationic pentanuclear clus-
ters, European Journal of Inorganic Chemistry, no. 10, pp.
19641978, 2005.
[28] T. C. Stamatatos, S. Dionyssopoulou, G. Efthymiou et al., The
rst cobalt metallacrowns: preparation and characterization
of mixed-valence cobalt(II/III), inverse 12-metallacrown-4
complexes, Inorganic Chemistry, vol. 44, no. 10, pp. 3374
3376, 2005.
[29] T. C. Stamatatos, A. K. Boudalis, Y. Sanakis, and C. P.
Raptopoulou, Reactivity and structural and physical studies
of tetranuclear iron(III) clusters containing the [Fe(-O)]
buttery core: an Fe cluster with an S = 1 ground state,
Inorganic Chemistry, vol. 45, no. 18, pp. 73727381, 2006.
[30] T. C. Stamatatos, J. C. Vlahopoulou, Y. Sanakis et al., For-
mation of the {Cu
II
3
(
3
-OH)}
5+
core in copper(II) carboxylate
chemistry via use of di-2-pyridyl ketone oxime [(py)
2
CNOH]
: [Cu
3
(OH)(O
2
CR)
2
{(py)
2
CNO}
3
] (R = Me, Ph), Inorganic
Chemistry Communications, vol. 9, no. 8, pp. 814818, 2006.
[31] C. Papatriantafyllopoulou, C. P. Raptopoulou, A. Escuer, and
C. J. Milios, A rare all-Mn decametallic cage from distorted
face-sharing cubes, Inorganica Chimica Acta, vol. 360, no. 1,
pp. 6168, 2007.
[32] K. V. Pringouri, C. P. Raptopoulou, A. Escuer, and T.
C. Stamatatos, Initial use of di-2-pyridyl ketone oxime
in chromium carboxylate chemistry: triangular Cr ( -O)
compounds and unexpected formation of a carboxylate-free
dichromium(II,II) complex, Inorganica Chimica Acta, vol.
360, no. 1, pp. 6983, 2007.
[33] T. C. Stamatatos, D. Foguet-Albiol, C. C. Stoumpos et al.,
New Mn structural motifs in manganese single-molecule
magnetism from the use of 2-pyridyloximate ligands, Poly-
hedron, vol. 26, no. 911, pp. 21652168, 2007.
[34] T. C. Stamatatos, E. Diamantopoulou, C. P. Raptopoulou,
V. Psycharis, A. Escuer, and S. P. Perlepes, Acetate/di-2-
pyridyl ketone oximate blend as a source of high-nuclearity
nickel(II) clusters: dependence of the nuclearity on the nature
of the inorganic anion present, Inorganic Chemistry, vol. 46,
no. 7, pp. 23502352, 2007.
[35] T. C. Stamatatos, C. Papatriantafyllopoulou, E. Katsoulakou,
C. P. Raptopoulou, and S. P. Perlepes, 2-Pyridyloximate
clusters of cobalt and nickel, Polyhedron, vol. 26, no. 911,
pp. 18301834, 2007.
[36] T. C. Stamatatos, A. Escuer, K. A. Abboud, C. P. Raptopoulou,
S. P. Perlepes, and G. Christou, Unusual structural types in
nickel cluster chemistry from the use of pyridyl oximes: Ni,
NiNa, and Ni clusters, Inorganic Chemistry, vol. 47, no. 24,
pp. 1182511838, 2008.
[37] M. Murugesu, K. A. Abboud, and G. Christou, New hexanu-
clear and dodecanuclear Fe(III) clusters with carboxylate and
alkoxide-based ligands from cluster aggregation reactions,
Polyhedron, vol. 23, no. 17, pp. 27792788, 2004.
[38] T. Afrati, C. Dendrinou-Samara, C. P. Raptopoulou, A. Terzis,
V. Tangoulis, and D. P. Kessissoglou, A tetranuclear mixed-
valent MnMn compound with a (-O)Mn core, Angewandte
ChemieInternational Edition, vol. 41, no. 12, pp. 21482150,
2002.
[39] A. J. Stemmler, J. W. Kampf, and V. L. Pecoraro, Synthesis
and crystal structure of the rst inverse 12-metallacrown-4,
Inorganic Chemistry, vol. 34, no. 9, pp. 22712272, 1995.
[40] T. C. Stamatatos, D. Foguet-Albiol, C. C. Stoumpos et al.,
Initial example of a triangular single-molecule magnet from
ligand-induced structural distortion of a [MnO] complex,
Journal of the American Chemical Society, vol. 127, no. 44, pp.
1538015381, 2005.
[41] C. Papatriantafyllopoulou, E. Manessi-Zoupa, A. Escuer, and
S. P. Perlepes, The sulfate ligand as a promising player in
3d-metal cluster chemistry, Inorganica Chimica Acta, vol. 362,
no. 3, pp. 634650, 2009.
[42] M. I. Khan, S. Cevik, and R. J. Doedens, Inorganic-organic
hybrids derived from oxovanadium sulfate motifs: synthesis
and characterization of [VO(-SO)(2,2-bpy)], Chemical
Communications, no. 19, pp. 19301931, 2001.
8 International Journal of Inorganic Chemistry
[43] A. C. Sudik, A. R. Millward, N. W. Ockwig, A. P. C ot e, J.
Kim, and O. M. Yaghi, Design, synthesis, structure, and gas
(N, Ar, CO, CH, and H) sorption properties of porous metal-
organic tetrahedral and heterocuboidal polyhedra, Journal of
the American Chemical Society, vol. 127, no. 19, pp. 71107118,
2005.
[44] J. Moriyama, H. Nishiguchi, T. Ishihara, and Y. Takita, Metal
sulfate catalyst for CCl
2
F
2
decomposition in the presence of
H
2
O, Industrial and Engineering Chemistry Research, vol. 41,
no. 1, pp. 3236, 2002.
[45] S. Y. Wan, YI. Z. Li, T. A. Okamura, J. Fan, W. Y.
Sun, and N. Ueyama, Syntheses, structures, and prop-
erties of two-dimensional honeycomb networks from the
assembly of the tripodal ligand 2,4,6-Tris[4-(imidazol-1-
ylmethyl)phenyl]-1,3,5-triazine with metal salts, European
Journal of Inorganic Chemistry, no. 20, pp. 37833789, 2003.
[46] Z. He, E. N. Q. Gao, Z. M. Wang, C. H. Yan, and M. Kurmoo,
Coordination polymers based on inorganic lanthanide(III)
sulfate skeletons and an organic isonicotinate N-oxide connec-
tor: segregation into three structural types by the lanthanide
contraction eect, Inorganic Chemistry, vol. 44, no. 4, pp.
862874, 2005.
[47] J. Reedijk, The relevance of hydrogen bonding in the
mechanism of action of platinum antitumor compounds,
Inorganica Chimica Acta, vol. 198200, pp. 873881, 1992.
[48] D. Peak, R. G. Ford, and D. L. Sparks, An in situ ATR-FTIR
investigation of sulfate bonding mechanisms on goethite,
Journal of Colloid and Interface Science, vol. 218, no. 1, pp. 289
299, 1999.
[49] G. Tamasi and R. Cini, Study of binary and ternary metal
complexes containing the sulfato ligand: molecular models for
selected non-catalytic sites in sulfurylase, Dalton Transactions,
no. 14, pp. 29282936, 2003.
[50] C. Papatriantafyllopoulou, L. F. Jones, T. D. Nguyen et al.,
Using pyridine amidoximes in 3d-metal cluster chemistry: a
novel ferromagnetic Ni complex from the use of pyridine-2-
amidoxime, Dalton Transactions, no. 24, pp. 31533155, 2008.
[51] C. Papatriantafyllopoulou, G. Aromi, A. J. Tasiopoulos et al.,
Use of the sulfato ligand in 3d-metal cluster chemistry: a
family of hexanuclear nickel(II) complexes with 2-pyridyl-
substituted oxime ligands, European Journal of Inorganic
Chemistry, no. 18, pp. 27612774, 2007.
[52] C. G. Efthymiou, A. A. Kitos, C. P. Raptopoulou, S. P. Perlepes,
A. Escuer, and C. Papatriantafyllopoulou, Employment of
the sulfate ligand in 3d-metal cluster chemistry: a novel
hexanuclear nickel(II) complex with a chair metal topology,
Polyhedron, vol. 28, no. 15, pp. 31773184, 2009.
[53] R. A. Coxall, S. G. Harris, D. K. Henderson, S. Parsons, P. A.
Tasker, and R. E. P. Winpenny, Inter-ligand reactions: in situ
formation of new polydentate ligands, Journal of the Chemical
Society, Dalton Transactions, no. 14, pp. 23492356, 2000.
[54] A. Altomare, G. Cascarano, C. Giacovazzo et al., SIR92a
program for automatic solution of crystal structures by direct
methods, Journal of Applied Crystallography, vol. 27, p. 435,
1993.
[55] G. M. Sheldrick, SHELXL-97, Program for the Renement
of Crystal Structures from Diraction Data, University of
G ottingen, G ottingen, Germany, 1997.
[56] P. Van der Sluis and A. L. Spek, BYPASS: an eective method
for the renement of crystal structures containing disordered
solvent regions, Foundations of Crystallography, vol. 46, no. 4,
pp. 194201, 1990.
[57] Oxford Diraction, CrysAlis CCD and CrysAlis RED, version
1.171.32.15; Oxford Diraction Ltd, Abingdon, UK, 2008.
[58] L. J. Farrugia, WinGX suite for small-molecule single-crystal
crystallography, Journal of Applied Crystallography, vol. 32,
no. 4, pp. 837838, 1999.
[59] I. J. Bruno, J. C. Cole, P. R. Edgington et al., New software for
searching the Cambridge Structural Database and visualizing
crystal structures, Acta Crystallographica Section B, vol. 58, no.
3, pp. 389397, 2002.
[60] K. Bradenburg, DIAMOND, Release 3.1f, Crystal Impact, GbR,
Bonn, Germany, 2008.
[61] V. L. Pecoraro, A. J. Stemmler, B. R. Gibney et al., Met-
allacrowns: a new class of molecular recognition agents,
Progress in Inorganic Chemistry, vol. 45, pp. 83177, 1997.
[62] C. Papatriantafyllopoulou, C. G. Efthymiou, C. P. Rap-
topoulou et al., Initial use of the di-2-pyridyl ketone/sulfate
blend in 3d-metal cluster chemistry: preparation, X-ray
structures and physical studies of zinc(II) and nickel(II)
cubanes, Journal of Molecular Structure, vol. 829, no. 13, pp.
176188, 2007.
[63] G. S. Papaefstathiou, A. Escuer, F. A. Mautner et al., Use
of the di-2-pyridyl ketone/acetate/dicyanamide blend in
manganese(II), cobalt(II) and nickel(II) chemistry: neutral
cubane complexes, European Journal of Inorganic Chemistry,
no. 5, pp. 879893, 2005.
[64] C. G. Efthymiou, C. P. Raptopoulou, A. Terzis et al., A sys-
tematic exploration of nickel(II)/acetate/di-2-pyridyl ketone
chemistry: neutral and cationic tetranuclear clusters, and a
novel mononuclear complex, European Journal of Inorganic
Chemistry, no. 11, pp. 22362252, 2006.
[65] C. Papatriantafyllopoulou, C. P. Raptopoulou, A. Terzis, J. F.
Janssens, S. P. Perlepes, and E. Manessi-Zoupa, Reactions of
nickel(II) sulfate hexahydrate with methyl(2-pyridyl)ketone
oxime: two mononuclear sulfato complexes containing the
neutral ligand, Zeitschrift fur Naturforschung, vol. 62, no. 9,
pp. 11231132, 2007.
[66] M. Moragues-C anovas, M. Helliwell, L. Ricard et al., An
Ni, single-molecule magnet: synthesis, structure and low-
temperature magnetic behavior, European Journal of Inor-
ganic Chemistry, no. 11, pp. 22192222, 2004.
[67] E. N. C. Yang, W. Wernsdorfer, S. Hill et al., Exchange bias in
Ni single-molecule magnets, Polyhedron, vol. 22, no. 1417,
pp. 17271733, 2003.
[68] A. Escuer, M. Font-Barda, S. B. Kumar, X. Solans, and R.
Vicente, Two new nickel(II) cubane compounds derived
from pyridine-2-methoxide (Pym): Ni(Pym)Cl(CHOH) and
Ni(Pym)(N)(CH OH). Crystal structures and magnetic prop-
erties, Polyhedron, vol. 18, no. 6, pp. 909914, 1999.
[69] E. N. C. Yang, W. Wernsdorfer, L. N. Zakharov et al., Fast
magnetization tunneling in tetranickel(II) single-molecule
magnets, Inorganic Chemistry, vol. 45, no. 2, pp. 529546,
2006.
[70] C. G. Efthymiou, C. P. Raptopoulou, A. Terzis et al., A sys-
tematic exploration of nickel(II)/acetate/di-2-pyridyl ketone
chemistry: neutral and cationic tetranuclear clusters, and a
novel mononuclear complex, European Journal of Inorganic
Chemistry, no. 11, pp. 22362252, 2006.
[71] C. G. Efthymiou, C. Papatriantafyllopoulou, N. I. Alex-
opoulou et al., A mononuclear complex and a cubane cluster
from the initial use of 2-(hydroxymethyl)pyridine in nickel(II)
carboxylate chemistry, Polyhedron, vol. 28, no. 15, pp. 3373
3381, 2009.
International Journal of Inorganic Chemistry 9
[72] A. Sieber, C. Boskovic, R. Bircher et al., Synthesis and spec-
troscopic characterization of a new family of Ni spin clusters,
Inorganic Chemistry, vol. 44, no. 12, pp. 43154325, 2005.
[73] K. Isele, F. Gigon, A. F. Williams et al., Synthesis, structure
and properties of {M
4
O
4
} cubanes containing nickel(II) and
cobalt(II), Dalton Transactions, no. 3, pp. 332341, 2007.
[74] M. A. Halcrow and G. Christou, Biomimetic chemistry of
nickel, Chemical Reviews, vol. 94, no. 8, pp. 24212481, 1994.
[75] A. J. Edwards, B. F. Hoskins, E. H. Kachab, A. Markiewicz, K.
S. Murray, and R. Robso, Synthesis, x-ray crystal structures,
and magnetic properties of Ni complexes of a macrocyclic
tetranucleating ligand, Inorganic Chemistry, vol. 31, no. 17,
pp. 35853591, 1992.
[76] E. Jabri, M. B. Carr, R. P. Hausinger, and P. A. Karplus, The
crystal structure of urease from Klebsiella aerogenes, Science,
vol. 268, no. 5213, pp. 9981004, 1995.
[77] A. Escuer, R. Vicente, S. B. Kumar, and F. A. Mautner, Spin
frustration in the buttery-like tetrameric [Ni(-CO)(aetpy)]-
[ClO ] [aetpy = (2-aminoethyl)pyridine] complex. Structure
and magnetic properties, Journal of the Chemical Society
Dalton Transactions, no. 20, pp. 34733477, 1998.
[78] J. M. Clemente, H. Andres, J. J. Borr as-Almenar et al.,
Magnetic excitations in polyoxometalate clusters observed
by inelastic neutron scattering: evidence for ferromagnetic
exchange interactions and spin anisotropy in the tetrameric
nickel(II) cluster [Ni(HO)(PWO)] and comparison with
the magnetic properties, Journal of the American Chemical
Society, vol. 121, no. 43, pp. 1002110027, 1999.
[79] Z. E. Serna, L. Lezama, M. K. Urtiaga et al., A dicubane-like
tetrameric nickel(II) azido complex, Angewandte Chemie
International Edition, vol. 39, no. 2, pp. 344347, 2000.
[80] P. Venkateswara Rao, S. Bhaduri, J. Jiang, and R. H. Holm,
Sulfur bridging interactions of cis-planar Ni-S N coor-
dination units with nickel(II), copper(I,II), zinc(II), and
mercury(II): a library of bridging modes, including Ni (-
SR)M rhombs, Inorganic Chemistry, vol. 43, no. 19, pp. 5833
5849, 2004.
[81] K. T. Szacilowski, P. Xie, A. Y. S. Malkhasian et al., Solid-state
structures and magnetic properties of halide-bridged, face-
to-face bis-nickel(II)-macrocyclic ligand complexes: ligand-
mediated interchanges of electronic conguration, Inorganic
Chemistry, vol. 44, no. 17, pp. 60196033, 2005.
[82] X. L opez, M. Y. Huang, G. C. Huang et al., Even-
numbered metal chain complexes: synthesis, characterization,
and DFT analysis of [Ni(-Tsdpda)(H O)] (Tsdpda = N-
(p-toluenesulfonyl) dipyridyldiamido), [Ni(-Tsdpda)] , and
related Ni string complexes, Inorganic Chemistry, vol. 45, no.
22, pp. 90759084, 2006.
[83] E. Carmona, E. Guti errez-Puebla, A. Monge, M. Paneque,
and M. L. Poveda, Unusual alkylidene-bridged complexes
of nickel by -H abstraction from a nickelacycle. Crystal
and molecular structure of [Ni (CHCMe-o-CH)Cl(PMe )],
Journal of the Chemical Society, Chemical Communications, no.
3, pp. 148150, 1991.
[84] J. Klingele, J. F. Boas, J. R. Pilbrow et al., A [2 2]
nickel(II) grid and a copper(II) square result from diering
binding modes of a pyrazine-based diamide ligand, Dalton
Transactions, no. 6, pp. 633645, 2007.
[85] F. Meyer, M. Konrad, and E. Kaifer, Novel -coordination of
urea at a nickel(II) site: structure, reactivity and ferromagnetic
superexchange, European Journal of Inorganic Chemistry, no.
11, pp. 18511854, 1999.
[86] K. Yamada, S. Yagishita, H. Tanaka et al., Metal-complex
assemblies constructed from the exible hinge-like ligand-
Hbhnq: structural versatility and dynamic behavior in the
solid state, ChemistryA European Journal, vol. 10, no. 11,
pp. 26472660, 2004.
[87] M. Bell, A. J. Edwards, B. F. Hoskins, E. H. Kachab, and R.
Robson, Synthesis and X-ray crystal structures of Ni and Zn
complexes of a macrocyclic tetranucleating ligand, Journal of
the American Chemical Society, vol. 111, no. 10, pp. 36033610,
1989.
[88] M. Alexiou, C. Dendrinou-Samara, C. P. Raptopoulou, A.
Terzis, V. Tangoulis, and D. P. Kessissoglou, A cationic
tetranuclear [Ni(MeOH) (pko)] cluster showing antiferro-
and ferromagnetic features, European Journal of Inorganic
Chemistry, no. 19, pp. 38223827, 2004.
[89] V. V. Pavlishchuk, S. V. Kolotilov, A. W. Addison et al.,
A tetrameric nickel(II) chair with both antiferromagnetic
internal coupling and ferromagnetic spin alignment, Ange-
wandte ChemieInternational Edition, vol. 40, no. 24, pp.
47344737, 2001.
[90] G. Psomas, C. Dendrinou-Samara, M. Alexiou et al., The
rst fused dimer metallacrown Ni
II
2
(mcpa)
2
(CH
3
OH)
3
(H
2
O)
[12-MC
Ni
II
N
(shi)2(pko)2
-4][12-MC
Ni
II
N
(shi)3(pko)
-4], Inorganic
Chemistry, vol. 37, no. 26, pp. 65566557, 1998.
[91] G. Psomas, A. J. Stemmler, C. Dendrinou-Samara et al.,
Preparation of site-dierentiated mixed ligand and mixed
ligand/mixed metal metallacrowns, Inorganic Chemistry, vol.
40, no. 7, pp. 15621570, 2001.
[92] M. Alexiou, I. Tsivikas, C. Dendrinou-Samara et al., High
nuclearity nickel compounds with three, four or ve metal
atoms showing antibacterial activity, Journal of Inorganic
Biochemistry, vol. 93, no. 3-4, pp. 256264, 2003.
[93] P. Chaudhuri, M. Winter, U. Fl orke, and H. J. Haupt, An
eectively diamagnetic oximato-bridged asymmetric dinu-
clear copper(II) complex with a Cu(II)I bond, Inorganica
Chimica Acta, vol. 232, no. 1-2, pp. 125130, 1995.
[94] C. Papatriantafyllopoulou, C. P. Raptopoulou, A. Terzis, E.
Manessi-Zoupa, and S. P. Perlepes, Investigation of the
zinc chloride/methyl(2-pyridyl)ketone oxime reaction system:
a mononuclear complex and an inverse 12-metallacrown-4
cluster, Zeitschrift fur Naturforschung, vol. 61, no. 1, pp. 37
46, 2006.
[95] J. J. L opez-Garriga, G. T. Babcock, and J. F. Harrison, Factors
inuencing the C=N stretching frequency in neutral and
protonated Schis bases, Journal of the American Chemical
Society, vol. 108, no. 23, pp. 72417251, 1986.
[96] A. B. P. Lever and E. Mantovani, The far-infrared and
electronic spectra of some bis-ethylenediamine and related
complexes of copper(II) and the relevance of these data to
tetragonal distortion and bond strengths, Inorganic Chem-
istry, vol. 10, no. 4, pp. 817826, 1971.
[97] K. Nakamoto, Infrared and Raman Spectra of Inorganic and
Coordination Compounds, John Wiley & Sons, New York, NY,
USA, 4th edition, 1986.
Hindawi Publishing Corporation
International Journal of Inorganic Chemistry
Volume 2011, Article ID 395418, 4 pages
doi:10.1155/2011/395418
Research Article
Improvement of Aminopeptidase Activity of Dizinc(II)
Complexes by Increasing Substrate Accessibility
Md. Jamil Hossain, Akinobu Wada, Yasuhiro Igarashi, Kei-ichiro Aimono, Keisuke Suzuki,
Katsuya Tone, and Hiroshi Sakiyama
Department of Material and Biological Chemistry, Faculty of Science, Yamagata University, Kojirakawa, Yamagata 990-8560, Japan
Correspondence should be addressed to Hiroshi Sakiyama, saki@sci.kj.yamagata-u.ac.jp
Received 22 November 2010; Accepted 2 March 2011
Academic Editor: Rabindranath Mukherjee
Copyright 2011 Md. Jamil Hossain et al. This is an open access article distributed under the Creative Commons Attribution
License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly
cited.
A new dizinc(II) complex, [Zn
2
(bhmp)(MeCO
2
)
2
]BPh
4
[(bhmp)

: 2,6-bis[bis(2-hydroxyethyl)aminomethyl]-4-methylphenolate
anion], performs aminopeptidase activity to hydrolyze L-leucine-p-nitroanilide. As compared with a related dizinc(II) complex
[Zn
2
(bomp)(MeCO
2
)
2
]BPh
4
[(bomp)

: 2,6-bis[bis(2-methoxyethyl)aminomethyl]-4-methylphenolate anion], the activity of the


present bhmp complex was about 80 times greater than that of the bomp complex. This is mainly because the substrate accessibility
was improved by changing the terminal methoxy groups to hydroxyl groups.
1. Introduction
Aminopeptidases are exopeptidases that remove the N-
terminal amino acid from a protein [15]. It is interesting
that most of the well-characterized aminopeptidases contain
dinuclear metal cores at their active sites, while carboxypepti-
dases, which remove the C-terminal amino acid, do not have
dinuclear cores. The well-characterized aminopeptidases are
leucine aminopeptidase (LAP, EC 3.4.11.1) [1], methionine
aminopeptidase (MAP, EC 3.4.11.18) [2], aminopeptidase
from Aeromonas proteolytica (AAP, EC 3.4.11.10) [3], Strep-
tomyces griseus aminopeptidase (SGAP, EC 3.4.11.-) [4],
and proline aminopeptidase (PAP, EC 3.4.11.9) [5]. LAP,
AAP, and SGAP contain dizinc(II) cores at their active sites,
while MAP contains a dicobalt(II) core and PAP contains a
dimanganese(II) core.
With the intention of nding a minimum functional unit
of aminopeptidase, Sakiyama and coworkers developed a
dizinc(II) complex which can be represented by the following
[Zn(bomp)(MeCO
2
)
2
]BPh
4
as the rst functional model of
aminopeptidase [6] represented by [(bomp)

: 2,6-bis[bis(2-
methoxyethyl)aminomethyl]-4-methylphenolate anion];
later, the aminopeptidase activity was improved by the intro-
duction of stronger electron-withdrawing p-substituents [7]
(the aminopeptidase activity was improved 10 times for the
chloro-substituted complex represented by the following
[Zn
2
(bocp)(MeCO
2
)
2
]BPh
4
and was improved 250 times
for the nitro-substituted complex represented by the
following [Zn
2
(bonp)(MeCO
2
)
2
]BPh
4
[(bocp)

: 4-chloro-
2,6-bis[bis(2-methoxyethyl)aminomethyl]phenolate anion;
(bonp)

: 2,6-bis[bis(2-methoxyethyl)aminomethyl]-4-ni-
trophenolate anion]). From the kinetic studies, the substrate
was proved to be incorporated within the dizinc center [7].
On the other hand, we found that the substrate acces-
sibility of the bomp complexes was not good because of
the steric hindrance of the terminal methoxy groups [7].
Therefore, in the present study, a new dizinc(II) complex,
[Zn
2
(bhmp)(MeCO
2
)
2
]BPh
4
(1), has been synthesized using
a dinucleating ligand, bhmp

[(bhmp)

: 2,6-bis[bis(2-
hydroxyethyl)aminomethyl]-4-methylphenolate anion] [8],
in which the methoxy groups of the bomp ligand are
substituted into less-hindered hydroxyl groups, and the
aminopeptidase activity of the complex was examined (see
Scheme 1).
2. Experimental
2.1. Measurements. Elemental analyses were obtained at the
Elemental Analysis Service Centre of Kyushu University.
2 International Journal of Inorganic Chemistry
N N
O O
O
R R
R R
O

O
CH
3
R = CH
3
(bomp

)
= H(bhmp

)
Scheme 1
Infrared (IR) spectra were recorded on a Hitachi 270-
50 spectrometer. Electronic spectra were recorded on a
Shimadzu UV-240 spectrophotometer.
2.2. Materials. Na(bhmp) was prepared as previously
described in [8]. All other chemicals were commercial
products and were used as supplied.
2.3. Synthesis of [Zn
2
(bhmp)(MeCO
2
)
2
]BPh
4
2MeCN (1).
To a methanolic solution (15 mL) of Na(bhmp) (0.19 g,
0.52 mmoL) was added zinc(II) acetate dihydrate (0.22 g,
1.00 mmoL), and the resulting solution was reuxed for
1 hour. The addition of sodium tetraphenylborate (0.17 g,
0.50 mmoL) resulted in the precipitation of colorless micro-
crystals, which were recrystallized from acetonitrile. Yield
0.26 g (52%). (Found: C, 59.15; H, 6.20; N, 5.65; Zn, 14.1.
Calc. for C
49
H
61
BN
4
O
9
Zn
2
: C, 59.35; H, 6.20; N, 5.65; Zn,
13.2%). Selected IR data [/cm
1
] using KBr disk: 3410,
3240, 30502850, 1605, 1580, 1475, 1420, 1330, 1260, 1130,
1030, 870, 730, 700, 605.
2.4. Aminopeptidase Activity. The aminopeptidase activity of
the complex was estimated using l-leucine-p-nitroanilide as
a substrate [6]. The substrate was dissolved in 1.5 mL of a
tricine buer solution (pH 8), and to this was added 1.0 mL
of a DMF solution of the complex at room temperature. The
hydrolysis of the substrate into l-leucine and p-nitroaniline
was monitored by detecting the formation of p-nitroaniline
using a spectrometer at 405 nm. In the measurement,
spontaneous hydrolysis of the substrate was subtracted as
a background. This measurement was examined at various
complex concentrations from 0 to 5.0 10
4
mol dm
3
and
at various substrate concentrations from 0 to 5.9 10
4
mol
dm
3
. This procedure was also carried out at pH values
3
2
1
0
0 2 4 6
[Complex]/10
4
(moldm
3
)
(
v
/
[
s
u
b
s
t
r
a
t
e
]
)
/
1
0

6
(
s

1
)
Figure 1: A v/[substrate] versus [complex] plot in the hydrolysis of
l-leucine-p-nitroanilide by 1 at nominal pH 8.
varying from 7 to 10 using HEPES (pH 7), tricine (pH 8),
and CHES (pH 9-10) buers.
3. Results and Discussion
3.1. Aminopeptidase Activity at pH 8. Before the discussion
about the aminopeptidase activity, it should be noted that
simple zinc salts, such as zinc(II) chloride and zinc(II)
sulfate, do not show aminopeptidase activity [6]. First, the
aminopeptidase activity of complex 1 was estimated in a
mixture of 40% DMF and 60% aqueous solution at pH 8
using l-leucine-p-nitroanilide as a substrate. Hereafter, the
term nominal pH will be used because the experiments
were carried out in a mixture of DMF and water. The mea-
surement was carried out at various complex concentrations
and at various substrate concentrations, and the initial rate
v was obtained for each experiment. A plot of the initial
rate over the substrate concentration v/[substrate] versus
the complex concentration [complex] showed good linearity
(Figure 1), as did the previous complexes [6, 7, 9], which
indicates that the initial rate can be written as a second-order
rate equation as follows:
v = k[substrate]
_
complex
_
, (1)
where k is the second-order rate constant. The k value
for the bhmp complex 1 was calculated as 4.4(2)
10
3
dm
3
mo1
1
s
1
. The previously obtained k value for the
bomp complex was 2.3(1) 10
3
dm
3
mo1
1
s
1
under the
same conditions [6]. Therefore, the rate for 1 was about two
times greater than that for the bomp complex at nominal pH
8.
3.2. Eect of pH on the Aminopeptidase Function. The above
procedure was also carried out at nominal pHs varying from
International Journal of Inorganic Chemistry 3
2.5
2
1.5
1
0.5
0
k
7 8 9 10 11
pH
Figure 2: k versus nominal pH plots for the bhmp complex 1 (O)
and the previous bomp complex ( ). The curves are drawn by (2)
using the parameters in the text.
7 to 10. At nominal pH 7, the activity was too small to
determine the rate constant, but in the nominal pH range
from 8 to 10, the second-order rate equation was found
to be valid. The k versus nominal pH plot for 1 is shown
in Figure 2. The plot is sigmoidal around the nominal pH
9.5, and the data could be tted using (2) with parameters
k
depro
= 2.55 dm
3
mo1
1
s
1
and pK
a
= 9.44, where k
depro
is
the rate constant for the deprotonated form.
k = k
depro
K
a
K
a
+ [H
+
]
. (2)
The result indicates that the reaction was promoted
by the deprotonated form of the complex, which will be
discussed in Section 3.3. In the case of the previous bomp
complex, the data was reexamined using (2), and the param-
eters were determined as k
depro
= 3.26 10
2
dm
3
mo1
1
s
1
and pK
a
= 9.07. The activity (k
depro
) of the present bhmp
complex 1 was about 80 times greater than that of the bomp
complex although the k value at nominal pH 8 was about
only 2 times greater. This is because the pKa value of 1 is
slightly larger than that of the bomp complex.
3.3. Some Considerations about the Active Species. Accord-
ing to the crystal structure of related cobalt(II) complex
[Co
2
(bhmp)(MeCO
2
)
2
]BPh
4
[8], the coordination geome-
try around each zinc(II) ion is saturated, and the two zinc(II)
ions are bridged by two acetate ions. However, the dissocia-
tion of the acetate ions occurs rather easily in an aqueous
solution, aording vacant coordination sites for substrate
incorporation. Indeed, in the cases of related cobalt(II)
and nickel(II) complexes, [Co
2
(bhmp)(MeCO
2
)
2
]ClO
4
and
[Ni
2
(bhmp)(MeCO
2
)
2
]ClO
4
, the dominant species in an
aqueous solution were identied as [M
2
(bhmp)(H
2
O)
4
]
2+
and [M
2
(bhmp)(MeCO
2
)(H
2
O)
2
]
+
(M = Co
II
, Ni
II
) by
analyzing the electronic spectra [10]. When one or two water
molecules of [Zn
2
(bhmp)(H
2
O)
4
]
2+
are exchanged with the
substrate and a remaining water molecule is deprotonated, a
nucleophilic Zn-OH moiety is thought to attack the carbonyl
carbon of the bound substrate.
4. Conclusion
For the purpose of improving the substrate accessibility, a
new dizinc(II) complex, [Zn
2
(bhmp)(MeCO
2
)
2
]BPh
4
(1),
was synthesized, and its aminopeptidase activity was inves-
tigated. As compared with the previous dizinc(II) complex,
[Zn
2
(bomp)(MeCO
2
)
2
]BPh
4
, the activity of 1 was about 80
times greater. Routine kinetic results have been deposited as
supplementary material at 10.1155/2011/395418.
References
[1] S. K. Burley, P. R. David, A. Taylor, and W. N. Lipscomb,
Molecular structure of leucine aminopeptidase at 2.7-

A
resolution, Proceedings of the National Academy of Sciences of
the United States of America, vol. 87, no. 17, pp. 68786882,
1990.
[2] S. L. Roderick and B. W. Matthews, Structure of the cobalt-
dependent methionine aminopeptidase from escherichia coli:
a new type of proteolytic enzyme, Biochemistry, vol. 32, no.
15, pp. 39073912, 1993.
[3] B. Chevrier, C. Schalk, H. DOrchymont, J. M. Rondeau,
D. Moras, and C. Tarnus, Crystal structure of Aeromonas
proteolytica aminopeptidase: a prototypical member of the
co-catalytic zinc enzyme family, Structure, vol. 2, no. 4, pp.
283291, 1994.
[4] H. M. Greenblatt, O. Almog, B. Maras et al., Streptomyces
griseus aminopeptidase: X-ray crystallographic structure at
1.75

A resolution, Journal of Molecular Biology, vol. 265, no.
5, pp. 620636, 1997.
[5] M. C. J. Wilce, C. S. Bond, N. E. Dixon et al., Structure
and mechanism of a proline-specic aminopeptidase from
Escherichia coli, Proceedings of the National Academy of
Sciences of the United States of America, vol. 95, no. 7, pp. 3472
3477, 1998.
[6] H. Sakiyama, R. Mochizuki, A. Sugawara, M. Sakamoto,
Y. Nishida, and M. Yamasaki, Dinuclear zinc(II) complex
of a new acyclic phenol-based dinucleating ligand with
four methoxyethyl chelating arms: rst dizinc model with
aminopeptidase function, Journal of the Chemical Society,
Dalton Transactions, no. 6, pp. 9971000, 1999.
[7] H. Sakiyama, Y. Igarashi, Y. Nakayama, M. J. Hossain,
K. Unoura, and Y. Nishida, Aminopeptidase function of
dinuclear zinc(II) complexes of phenol-based dinucleating
ligands: eect of p-substituents, Inorganica Chimica Acta, vol.
351, pp. 256260, 2003.
[8] M. J. Hossain, M. Yamasaki, M. Mikuriya, A. Kuribayashi, and
H. Sakiyama, Synthesis, structure, and magnetic properties
of dinuclear cobalt(II) complexes with a new phenol-based
dinucleating ligand with four hydroxyethyl chelating arms,
Inorganic Chemistry, vol. 41, no. 15, pp. 40584062, 2002.
4 International Journal of Inorganic Chemistry
[9] H. Sakiyama, K. Ono, T. Suzuki, K. Tone, T. Ueno, and
Y. Nishida, Aminopeptidase function of dinuclear zinc(II)
complexes with chiral dinucleating ligands: stereoselectivity by
chiral substrate recognition, Inorganic Chemistry Communi-
cations, vol. 8, no. 4, pp. 372374, 2005.
[10] A. Kazama, A. Wada, H. Sakiyama, M. J. Hossain, and Y.
Nishida, Synthesis of water-soluble dinuclear metal com-
plexes [metal = cobalt(II) and nickel(II)] and their behavior
in solution, Inorganica Chimica Acta, vol. 361, no. 9-10, pp.
29182922, 2008.
Hindawi Publishing Corporation
International Journal of Inorganic Chemistry
Volume 2011, Article ID 291567, 8 pages
doi:10.1155/2011/291567
Research Article
Synthesis and Crystal Structure Differences
between Fully and Partially Fluorinated
-Diketonate Metal (Co
2+
, Ni
2+
, and Cu
2+
) Complexes
Akiko Hori
1, 2
and Masaya Mizutani
1
1
Department of Chemistry, School of Science, Kitasato University, 1-15-1 Kitasato, Minami-ku, Sagamihara,
Kanagawa 252-0373, Japan
2
PRESTO, Japan Science and Technology Agency (JST), 4-1-8 Honcho, Kawaguchi, Saitama 332-0012, Japan
Correspondence should be addressed to Akiko Hori, hori@kitasato-u.ac.jp
Received 28 December 2010; Accepted 28 February 2011
Academic Editor: Daniel L. Reger
Copyright 2011 A. Hori and M. Mizutani. This is an open access article distributed under the Creative Commons Attribution
License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly
cited.
Coordination complexes, [Co
2
(1)
4
(H
2
O)
2
] (2), [Ni
2
(1)
4
(H
2
O)
2
] (3), and [Cu(1)
2
] (4), by using an asymmetric and partially
uorinated 3-hydroxy-3-pentauorophenyl-1-phenyl-2-propen-1-one (H1) have been prepared, and the structures were
investigated to compare with the corresponding fully uorinated complexes of [Co
2
(5)
4
(H
2
O)
2
] (6), [Ni
2
(5)
4
(H
2
O)
2
] (7),
and [Cu(5)
2
] (8) with bis(pentauorobenzoyl)methane (H5) and to understand the uorine-substituted eects. While the
coordination mode of the partially uorinated complexes was quite similar to the fully uorinated complexes, the intra- and
intermolecular -interactions of the ligand moieties were highly inuenced by the uorination eects; the arene-peruoroarene
interactions were observed in complexes 2 and 3 as a reason of the dinucleation. In this paper, we describe detail structures of the
protonated form of the ligand, H1, and complexes 24 by X-ray crystallographic studies.
1. Introduction
While - interactions are observed in various aromatic
compounds in chemistry and biology [17], the control of
the interaction is still challenging because of their repulsion
of the electron charge of the aromatic moieties. For example,
a benzene molecule has a negative quadrupole moment,
29.0 10
40
Cm
2
[8] stabilizing a rectangular orientation
of CH interaction [9] and a sliding orientation of
- stacking. The negative charge of aromatic center also
interacts the cationic source through cation inter-
actions [1012]. On the other hand, the electron charge
of aromatic compounds can be controlled by uorination
(e.g., a hexauorobenzene molecule which shows positive
quadrupole moment, 31.7 10
40
Cm
2
) [8] and unique
electrostatic interactions, such as arene-peruoroarene [13
18] and anion- [1921] interactions, which have attracted
interests in a couple of decades. We have also synthesized
several uorinated compounds to control the electrostatic
interactions and to design the metal metal [22] and
metal [23] arrangements through the interactions [24
28]. In these studies, the uorination into the aromatic
moieties on coordination complexes has potentially shown
three unique eects: (1) uorinated -planes show unique
molecular recognition through the electrostatic interactions;
(2) an electron-density of a metal is controlled by the uori-
nation; (3) uorinated -planes of the molecules are twisted
with the whole molecular planes by the steric hindrance
of the uorine substituents in several cases [23, 24]. Such
eects give rise to several unique guest-recognitions of the
coordination complexes based on the crystal engineering.
In this paper, we show the crystal structures of par-
tially uorinated H1 (Scheme 1) and its three coordination
complexes, [Co
2
(1)
4
(H
2
O)
2
] (2), [Ni
2
(1)
4
(H
2
O)
2
] (3), and
[Cu(1)
2
] (4) (Scheme 2). The corresponding coordination
complexes [M(DBM)
2
] (M = Co
2+
, Ni
2+
, Cu
2+
) with diben-
zoylmethanide ligand (DBM, C
6
H
5
COCHCOC
6
H
5

) [29
34] and [Co
2
(5)
4
(H
2
O)
2
] (6), [Ni
2
(5)
4
(H
2
O)
2
] (7), and
2 International Journal of Inorganic Chemistry
O F
F
F
F
F
DBM 1 5
O F
F
F
F
F
F
F
F
F
F O O

Scheme 1
O
O O
O
Cu
F
F
F F
F
F
F
F F
F
O O F
F
F
F
F
O
O
F
F
F
F
F
M
O O F
F
F
F
F
O
O
F
F
F
F
F
M
M(OAc)
2
4H
2
O
H
2
O
OH
2
Cu(OAc)
2
H
2
O
NaOMe
H1
4
2: M = Co, 3: M = Ni
Scheme 2
[Cu(5)
2
] (8) with fully uorinated ligand 5 [24] are also
discussed as a comparison.
2. Experimental
2.1. General. All chemicals were of reagent grade and used
without further purication. Fully uorinated compounds,
H5 and 68, were previously reported [24].
1
HNMRspectral
data were recorded on a Bruker DRX600 spectrometer. The
melting points were determined by a Yanako MP-500D
melting point apparatus. Infrared and electronic absorption
spectra were recorded on a Shimadzu IR 8400s and JASCO
V-660 spectrometer, respectively. The results of elemental
analysis of C and H were collected by PerkinElmer PE2400
analyzer.
2.2. Synthesis and Physical Properties of H1 and 24
3-hydroxy-3-pentauorophenyl-1-phenyl-2-propen-1-one (H1)
[35]. Dry benzene solutions of acetophenone (47 mmol)
and ethyl pentauorobenzoate (50 mmol) were slowly added
into NaH (120 mmol), and the mixture was reuxed for
10 h. Then, the mixture was added into 10% HClaq. and
extracted by AcOEt, dried by MgSO
4
, evaporated to remove
the solvent. After puried by column chromatography (silica
gel, 30% AcOEt/hexane), white powder of H1 was obtained
in 10% yield. Colorless needle crystals of H1 were obtained
from an ethanol solution, suitable for X-ray crystallographic
studies. Yield 10%. mp 116

C. EI-MS: m/z 314.


1
HNMR
(600 MHz, CDCl
3
) 15.89 (s, 1H), 7.94 (d, J = 7.2 Hz, 2H),
7.60 (t, J = 7.2 Hz, 1H), 7.50 (t, J = 7.2 Hz, 2H), 6.50 (s, 1H).
IR (KBr disk, cm
1
): 1653, 1576, 1520, 1490, 1261, 1193, 997,
988, 781. Elemental analysis: calcd. for C
15
H
7
F
5
O
2
(%): C
57.34, H 2.25; found: C 57.42, H 2.25.
[Co
2
(1)
4
(OH
2
)
2
] (2). A solution of H1 (0.25 g, 0.80 mmol)
in 1 : 1 ethanol-CH
2
Cl
2
solution (20 mL) was added to a
solution of Co(OAc)
2
4H
2
O (0.11 g, 0.44 mmol) in ethanol
(10 mL). The mixture was stirred for 2 h at r.t. and was evap-
orated to give reddish brown powder of 2. The powder was
extracted with CHCl
3
and crystallized in CH
2
Cl
2
/benzene
to give complex 22H
2
O3C
6
H
6
suitable for X-ray crys-
tallographic studies. The compound was dried by vacuum
conditions for measurement of physical properties. Yield
92%. mp 220

C. UV-Vis {CH
2
Cl
2
, nm ( M
1
cm
1
)}:
550 (100), 330 (76800), 254 (38200). IR (KBr disk, cm
1
):
3472(broad), 3328(broad), 3072, 1655, 1598, 1566, 1508,
1486, 1417, 1321, 1261, 1064, 988, 716. Elemental analysis:
International Journal of Inorganic Chemistry 3
C4
C5
C6
C7
C8
C15
C10
C11
C13
C12
C14
C9 C2
C3
C1
O1
F1
F2
F4
F3
F5
O2
(a)
(b)
Figure 1: (a) ORTEP drawing of the crystal structure with 50%
proba-bility thermal ellipsoids and (b) a part of packing structure
of H1 at 100 K (color scheme: C: gray; F: purple; O: red).
calcd. for C
66
H
34
Co
2
F
20
O
10
(2C
6
H
6
): C 53.39, H 2.31;
found: C 53.02, H 2.41.
[Ni
2
(1)
4
(OH
2
)
2
] (3). This was obtained as green powder
by the same procedure as 2 with Ni(OAc)
2
4H
2
O. The
single crystals of 32H
2
O3C
6
H
6
were obtained from a
CH
2
Cl
2
/benzene solution suitable for X-ray crystallographic
studies. The compound was dried by vacuum conditions
for measurement of physical properties. Yield 85%. mp
260

C. UV-Vis {CH
2
Cl
2
, nm ( M
1
cm
1
)}: 640 (20), 339
(62400), 253 sh (35200). IR (KBr disk, cm
1
): 3443(broad),
3336(broad), 3071, 1654, 1599, 1568, 1509, 1488, 1450, 1433,
1424, 1405, 1322, 1261, 1064, 988, 715. Elemental analysis:
calcd. for C
60
H
28
Ni
2
F
20
O
10
(3): C 51.25, H 2.01; found: C
51.38, H 2.33.
[Cu(1)
2
] (4). This was obtained as bluish green crystals by
the same procedure as 2 with CuCl
2
2H
2
O and NaOMe. The
single crystals of 4 were obtained from a CH
2
Cl
2
/benzene
solution suitable for X-ray crystallographic studies. Yield
97%. mp 319

C. IR (KBr disk, cm
1
): 1655, 1591, 1559,
1514, 1485, 1451, 1410, 1329, 1269, 986754, 710. Elemental
analysis: calcd. for C
30
H
12
CuF
10
O
4
: C 52.22, H1.75; found:
C 52.17, H 1.71.
2.3. Crystal Structure Determination. Single crystal X-ray
structures were determined on a Bruker SMART APEX CCD
diractometer with graphite monochromated MoK ( =
0.71073

A) generated at 50 kV and 35 mA. All crystals were
coated by paraton-N and were measured at 100 K.
Crystal data for H1: C
15
H
7
F
5
O
2
, Mw = 314.21, mon-
oclinic, P2
1
/c, a = 7.1675(10)

A, b = 7.1980(10)

A, c =
24.239(3)

A, = 93.086(2)

, V = 1248.7(3)

A
3
, Z = 4,

calcd
= 1.671 g cm
3
, GOF = 1.022, R((I) > 2(I)) =
0.0354, wR(F
o
2
) = 0.1027, CCDC 805820; 22H
2
O3C
6
H
6
:
C
78
H
50
Co
2
F
20
O
12
, Mw = 1677.04, triclinic, P-1, a =
9.7381(5)

A, b = 13.7690(7)

A, c = 14.5108(7)

A, =
114.060(1)

, = 93.870(1)

, = 96.571(1)

, V =
1750.92(15)

A
3
, Z = 1,
calcd
= 1.590 g cm
3
, GOF = 1.097,
R((I) > 2(I)) = 0.0285, wR(F
o
2
) = 0.0852, CCDC
805821; 32H
2
O3C
6
H
6
: C
78
H
50
F
20
Ni
2
O
12
, Mw = 1676.60,
triclinic, P-1, a = 9.7085(9)

A, b = 13.7652(13)

A, c =
14.5171(13)

A, = 114.385(1)

, = 93.688(1)

, =
96.484(1)

, V = 1742.4(3)

A
3
, Z = 1,
calcd
=
1.598 g cm
3
, GOF = 1.052, R((I) > 2(I)) = 0.0296,
wR(F
o
2
) = 0.0773, CCDC 805822; 4: C
30
H
12
CuF
10
O
4
,
Mw = 689.94, monoclinic, P2
1
/c, a = 11.958(5)

A, b =
6.273(3)

A, c = 17.310(8)

A, = 107.549(5)

, V =
1237.9(10)

A
3
, Z = 2,
calcd
= 1.851 g cm
3
, GOF =
0.998, R((I) > 2(I)) = 0.0387, wR(F
o
2
) = 0.0939,
CCDC 805823. These data can be obtained free of charge
from The Cambridge Crystallographic Data Center via
http://www.ccdc.cam.ac.uk/data request/cif/.
3. Results and Discussion
3.1. Preparation and Structure of H1. The protonated form
of the ligand, H1, was prepared by previously reported
procedure [35] and characterized by
1
HNMR, EI-MS, and
elemental analysis. Colorless needle crystals of H1 were
obtained from an ethanol solution, suitable for X-ray crystal-
lography. The ORTEP drawing and a part of packing struc-
tures of crystal H1 are shown in Figure 1. The OHproton was
located into the side of the pentauorophenyl group, based
on the dierence Fourier density map and rened as riding
on its idealized position, O2-H = 0.84

A. The bond distances
of the O1-C7 and O2-C9 are 1.2841(18) and 1.2918(19)

A,
respectively. The r.m.s deviation of the -diketonato plane,
O1-C7-C8-C9-O2, is 0.008

A. The pentauorophenyl group
is more twisted to the plane of O1-C7-C8-C9-O2 than the
phenyl group; the dihedral angle between O1-C7-C8-C9-
O2 and the pentauorophenyl group of C10-C11-C12-C13-
C14-C15 is 41.42(4)

and that between O1-C7-C8-C9-O2


and the phenyl group of C1-C2-C3-C4-C5-C6 is 16.23(6)

.
Interestingly, the H1 shows head-to-tail stacking through
arene-peruoroarene interactions to give an alternate layered
structure; the closest intermolecular distance and the cor-
responding perpendicular distance of CgPh CgC
6
F
5

i
(i: x + 1, y 0.5, z +0.5) are 3.6838(10) and 3.4529(6)

A,
respectively, where CgPh is the centroid of the phenyl
ring and CgC
6
F
5
is the centroid of the pentauorophenyl
ring. The carbonyl moiety also closely interacts to pentau-
orophenyl group, and the intermolecular distance of C7-
O1 CgC
6
F
5

ii
(ii: x+1, y+0.5, z+0.5) is 3.2981(14)

A.
This -stacked structure including the interaction of phenyl
and pentauorophenyl groups prompted us to investigate of
the synthesis and crystallographic studies for its coordination
complexes.
3.2. Preparation and Structures of Co
2+
and Ni
2+
Complexes.
Co
2+
and Ni
2+
complexes with the partially uorinated
4 International Journal of Inorganic Chemistry
C37
C38
C28
C27
C29
C30
C25
C24
C26
C23
C22
C19
C18
C17
C16
C11
C12
C13
C14
C15
C5 C10
C9
C8
C4
C7
C6
C3
C2
C1
C31
C32
C33
C34
C36
C35
Co
O4
O6
O5
O1
O3
O2
C20
C21
F8
F9
F10
F7
F6
F1
F5
F4
F3
F2
C39
Figure 2: Numbering scheme of an asymmetric unit of the crystal structure of 22H
2
O3C
6
H
6
at 100 K. ORTEP drawing is shown with
50% probability thermal ellipsoids.
ligand 1 were obtained as dinuclear complexes (Scheme 2),
which are the same tendency of the fully uorinated
complexes of 6 and 7 [24]. Typically, M(OAc)
2
4H
2
O (M
= Co, Ni) and H1 were combined in an ethanol/CH
2
Cl
2
solution to give [Co
2
(1)
4
(OH
2
)
2
] 2 and [Ni
2
(1)
4
(OH
2
)
2
] 3.
These complexes were crystallized from CH
2
Cl
2
with vapor-
phase diusion of benzene to give red block crystals of
22H
2
O3C
6
H
6
and green block crystals of 32H
2
O3C
6
H
6
.
Fundamentally, the reaction of DBM and Co
2+
/Ni
2+
ions
formed mononuclear complexes, [M(DBM)
2
(X)
2
] (M =
Co
2+
or Ni
2+
, X = solvent or water), and the DBM ligands
were sited in equatorial planes [30]. Thus, the dinucleation
of the complexes with partially and fully uorinated ligands,
1 and 5, are quite unique motifs, which are useful to under-
stand the intramolecular interaction of the coordination
complexes.
The crystal structure of 22H
2
O3C
6
H
6
is shown in
Figures 2 and 3(a), and 22H
2
O3C
6
H
6
and 32H
2
O3C
6
H
6
are isomorphs. The selected bond distances and angles
of 22H
2
O3C
6
H
6
, 32H
2
O3C
6
H
6
, 62C
6
H
6
, and
72C
6
H
6
are summarized in Table 1. The detail structure of
22H
2
O3C
6
H
6
is as follows. Complex 2 comprises two Co
2+
ions, four ligands 1, and two water molecules to give the
dinuclear complex. The complex lies across crystallographic
inversion center. Both of the geometries around the metal
centers are pseudo-octahedral. The two ligands are chelate
coordinated to each metal by the O1 and O2 (x, y, z) and
O1
iii
and O2
iii
(iii: x +2, y + 1, z + 1) atoms. Two metal
centers are linked through the O4 and O4
iii
atoms forming a
lozenge geometry of the dinuclear core, and the remaining
O3 and O3
iii
atoms are coordinated to each metal. The O5
and O5
iii
atoms of two water molecules are coordinated to
each metal and two water solvates, O6 and O6
iii
, and linked
to the coordinated water through hydrogen bonds; the
intermolecular distance of O5 O6 is 2.6888(19)

A and the
angle of O5-H40B-O6 is 174(2)

, as shown in Figure 2. The


metal metal separations are 3.2061(3)

A for 2. The M
O(ligand) distances are 2.0265(10), 2.0159(10), 2.0422(10),
2.0529(10), and 2.1492(10)

A for 2 (av. 2.06

A). The
CoO5(water) distances of 2 is 2.0755(11), and the average
of the O=C bond distances is 1.27

A. The r.m.s deviations
of O1-C7-C8-C9-O2 and O3-C22-C23-C24-O4 in 2 are
0.0231 and 0.0112

A, respectively. The pentauorophenyl
groups, C10-C11-C12-C13-C14-C15 and C25-C26-C27-
C28-C29-C30, of 2 are highly twisted to the coordination
plane; the dihedral angles between O1-C7-C8-C9-O2 and
C10-C11-C12-C13-C14-C15 is 36.75(5)

and that between


O1-C7-C8-C9-O2 and C25-C26-C27-C28-C29-C30 is
82.47(5)

. The dihedral angles between O1-C7-C8-C9-O2


and two phenyl groups, C1-C2-C3-C4-C5-C6 and C16-C17-
C18-C19-C20-C21, are 17.19(5)

and 18.93(9)

, respectively.
The dihedral angle for C25-C26-C27-C28-C29-C30 is
remarkably twisted because of the intramolecular stacking of
the phenyl group of C1
iii
-C2
iii
-C3
iii
-C4
iii
-C5
iii
-C6
iii
through
the arene-peruoroarene interaction; the intramolecular
distance and the corresponding perpendicular distance of
CgC
6
F
5
CgPh
iii
are 4.3461(10) and 3.3637(7)

A,
respectively. This intramolecular stacking leading to
the ecient overlapping of the ligands gives ecient
stabilization of the dinuclear framework (Figure 3(a)).
The similar structure was observed in Ni
2+
complex,
32H
2
O3C
6
H
6
. Complex 3 also comprises two Ni
2+
ions,
International Journal of Inorganic Chemistry 5
Table 1: Selected bond distances and angles of Co and Ni complexes.
Complex 22H
2
O3C
6
H
6
32H
2
O3C
6
H
6
62C
6
H
6
2 [24] 72C
6
H
6
[24]
M M
iii/iv
3.2061(3)

A 3.1638(4) 3.1822(3) 3.1385(3)
M-O1 2.0265(10) 1.9891(10) 2.0132(11) 1.9842(11)
M-O2 2.0159(10) 1.9952(10) 2.0407(11) 2.0138(12)
M-O3 2.0422(10) 2.0051(10) 2.0496(11) 2.0087(12)
M-O4 2.0529(10) 2.0348(10) 2.0465(11) 2.0231(11)
M-O4
iii/iv
2.1492(10) 2.1101(10) 2.1628(11) 2.1224(12)
M-O5 2.0755(11) 2.0567(11) 2.1127(13) 2.0828(13)
O1-C7 1.2685(17) 1.2687(17) 1.2616(19) 1.266(2)
O2-C9 1.2774(18) 1.2739(18) 1.2722(19) 1.271(2)
O3-C22 1.2596(17) 1.2573(18) 1.2506(19) 1.252(2)
O4-C24 1.2850(17) 1.2831(17) 1.2931(19) 1.293(2)
O1-M-O2 89.23(4) 91.13(4) 89.61(4) 91.74(5)
O3-M-O4 88.45(4) 90.01(4) 89.27(4) 90.79(5)
O4-M-O4
iii/iv
80.57(4) 80.51(4) 81.82(4) 81.61(5)
M-O4-M
iii/iv
99.43(4) 99.49(4) 98.18(4) 98.39(5)
O2-M-O5 169.54(4) 171.49(4) 171.63(5) 172.70(5)
four 1, and two water molecules. Both of the geometries
around the metal centers are pseudo-octahedral. The coor-
dination bond distances between the oxygen atoms and Ni
2+
are also summarized in Table 1 and the average M-O(ligand)
distance of 32H
2
O3C
6
H
6
is slightly shorter than that
of 22H
2
O3C
6
H
6
, as depends on the size of metal ions.
The pentauorophenyl group of 3 interacts phenyl group
through the same intramolecular interactions.
While the molecular structures of the complexes 2 and
3 are resembled to the corresponding complexes 6 and 7
(see Figure 3), the crystal solvates are dierent; two water
molecules and three benzene molecules are included for the
partially uorinated complexes and two benzene molecules
are included for the fully uorinated complexes [24]. The
structures of 2 and 6 are discussed as follows. The average
of MO(ligand) distances is the same, 2.06

A, for 2 and 6,
and the average of the O=C bond distances is the same,
1.27

A, for 2 and 6. The CoO5(water) distances of 2
[2.0755(11)

A] are shorter than those of 6 [2.1127(13)

A].
The metal metal separation of 2 [3.2061(3)

A] is slightly
longer than that of 6 [3.1822(3)

A]. The Ni
2+
complexes 3
and 7 have also the same dierence; the average O=C bond
distances and metal metal separation of 3 are shorter
and longer, respectively, than those of 7. It is pointed
out that the pentauorophenyl groups of 6 (and 7) are
also highly twisted with respect to the coordination plane
(the torsion angles C5-C6-C7-C8, C8-C9-C10-C15, C20-
C21-C22-C23, and C23-C24-C25-C30 are 38.9(2), 63.6(2),
35.7(2), and 68.2(2)

, resp.) and the intramolecular stacking


is dominant between the two rings, C25-C26-C27-C28-C29-
C30 and C1
iv
-C2
iv
-C3
iv
-C4
iv
-C5
iv
-C6
iv
(iv: x + 1, y, z);
the intramolecular distance and the corresponding perpen-
dicular distance of CgC
6
F
5
CgC
6
F
5

iv
are 4.4373(10)
and 3.3253(7)

A, respectively. The pentauorophenyl groups
have a twisted conformation with the coordination plane,
leading to the ecient overlapping of the ligands for
dinuclear complexes (Figure 3(b)). This feature is in contrast
to the case of the complexes of DBM [30], where phenyl
groups and the coordination plane are essentially planar
due to the expanding -conjugation, which causes a steric
hindrance and hence mononucleation.
In the packing structures of 22H
2
O3C
6
H
6
and
32H
2
O3C
6
H
6
, three benzene molecules (two C31-C32-
C33-C34-C35-C36 and one C37-38-C39-C37
v
-C38
v
-C39
v
(v: x + 2, y + 2, z)) are included in the crystals and
these benzenes interact with the pentauorophenyl groups
through arene-peruoroarene interactions. This capsulated
solvate through the interactions was also observed in the
fully uorinated complexes of 62C
6
H
6
and 72C
6
H
6
[24].
Especially, the crystals of 7 show unique pseudopolymorphs
of 72C
6
H
6
and 74C
6
H
6
[25]. In our examinations, when
no benzene solvent was used in the crystallization, single
crystals were not grown as suitable for X-ray crystallography.
These results indicate that the crystals of the partially and
fully uorinated Co
2+
and Ni
2+
complexes are stabilized by
benzene molecules through arene-peruoroarene interac-
tions.
3.3. Preparation and Structure of Cu
2+
Complex. The Cu
2+
complex with the ligand 1 was obtained as a mononu-
clear complex (Scheme 2). Typically, an MeOH (10 mL)
solution of H1 (0.10 g, 0.30 mmol) and NaOMe (16 mg,
0.3 mmol) was combined into an MeOH (5 mL) solution
of CuCl
2
2H
2
O (30 mg, 0.15 mmol). Then, the mixture was
stirred for 2 h at r.t. to give a green precipitates of [Cu(1)
2
] 4.
In this case, the structure of 4 is mononuclear complex and
no inuences are observed by uorine substitutions because
the DBM and H5 are also mononuclear complexes [24]. This
result is expected from the fact that the ligands are only
6 International Journal of Inorganic Chemistry
O1
O5
O2
O3
O4
Co
Co
iii
(a)
O1
O5
O2
O3
O4
Co
Co
iv
(b)
Figure 3: ORTEP drawings of the dinuclear complexes of (a)
22H
2
O3C
6
H
6
and (b) 62C
6
H
6
at 100 K with 50% probability
thermal ellipsoids. For (a) and (b), symmetry transformations used
to generate equivalent atoms show iii (x + 2, y + 1, z + 1) and
iv (x + 1, y, z), respectively.
sited in equatorial planes of Cu
2+
ions by Jahn-Teller eect.
The Crystallization by diusion of benzene into a CH
2
Cl
2
solution of 4 yielded the pure product. The crystallization of
three Cu
2+
complexes, the target complex 4, [Cu(DBM)
2
],
and the fully uorinated complex 8 in CH
2
Cl
2
/benzene
conditions gave single crystals, 4, [Cu(DBM)
2
], and 83C
6
H
6
[24]. While crystals of the fully uorinated Cu
2+
complex
capsulated 21 w% of benzene molecules, no solvate crystals
were obtained for partially uorinated Cu
2+
complex 4, as
well as [Cu(DBM)
2
].
In the crystal, complex 4 comprises one Cu
2+
and two lig-
and 1 (Figure 4), that the composition and the packing struc-
ture was similar to [Cu(DBM)
2
]. The complex lies across
crystallographic inversion center. The geometry around the
metal is essentially planar, forming a square-planar. The r.m.s
deviation of O1-C7-C8-C9-O2 is 0.016

A showing the at
plane of the chelate moiety of the -diketonate framework.
The bond distances of the Cu-O1, Cu-O2, O1-C7, and O2-
C9 are 1.9100(19), 1.9216(18), 1.271(3), and 1.271(3)

A,
O2
F1
F5
F4
F3
F2
O1
Cu
C1
C2
C3
C4
C5
C6
C7
C8
C9
C10
C15
C14
C11
C12
C13
O2
vi
O1
vi
Figure 4: ORTEP drawings of the crystal structure of 4 with 50%
probability thermal ellipsoids. Symmetry transformation used to
generate equivalent atoms is vi (x, y + 1, z).
(a)
(b)
Figure 5: Parts of packing structures of (a) 4 and (b) [Cu(DBM)
2
].
respectively. The phenyl and pentauorophenyl groups are
clearly dierent orientation in 4. The phenyl group is
at with respect to the coordination plane (the dihedral
angle between O1-C7-C8-C9-O2 and C1-C2-C3-C4-C5-C6
is 6.82(16)

), and the pentauorophenyl group is highly


twisted (that between O1-C7-C8-C9-O2 and C10-C11-C12-
C13-C14-C15 is 56.76(8)

). On the other hand, the crystal


structure of [Cu(DBM)
2
] at 293 K is highly at [31, 32]. The
phenyl groups of [Cu(DBM)
2
] are also at with respect to the
coordination plane, and the dihedral angles between O1-C7-
C8-C9-O2 and two phenyl rings are 9.96(30) and 4.98(33)

,
International Journal of Inorganic Chemistry 7
showing the -delocalization of the whole complex. The
averages of the Cu-O and O=C bond distances are 1.90

A
and 1.27

A, respectively. The fully uorinated Cu
2+
complex
8 was obtained as 83C
6
H
6
under the same crystallization
procedure of CH
2
Cl
2
/benzene [23, 24]. The averages of the
Cu-O(5) and O=C bond distances of 8 are 1.92

A and 1.27

A,
respectively. The pentauorophenyl groups of 8 are also
highly twisted with respect to the coordination plane [24],
giving the benzene capsulated cavities.
Parts of the crystal packing of 4 and [Cu(DBM)
2
] were
shown in Figures 5(a) and 5(b), respectively. A common
intermolecular electrostatic interaction is found in both
crystals, such as a cation- interaction. The phenyl groups
in 4 interact with Cu
2+
ion and the pentauorophenyl
groups of another molecule through the cation- and
arene-peruoroarene interactions, respectively. The closest
intermolecular distances of CgPh Cu
vii
(vii: x, y + 1, z)
and CgPh CgC
6
F
5

viii
(viii: x, y + 0.5, z + 0.5) are
3.270

A and 3.631

A, respectively. This cation- interaction
is only observed in Cu
2+
complexes because of the free
apical sites of the Cu
2+
ions. The CF interaction is
also observed between the two pentauorophenyl groups,
and the distance of F4 atom in the pentauorophenyl group
and CgC
6
F
5

ii
is 3.109

A, which is a reversed version of the
charge orientation of the CH interaction.
In crystal of 8, one benzene molecule closely interacts
with the Cu
2+
ion through the cation- interaction, and
two benzenes interact with the pentauorophenyl groups
through the arene-peruoroarene interaction. This result
indicates, that the Cu
2+
ion preferably recognized -molecu-
les and the cation- interaction is required for the molecular
packing. The crystallographic study of 4 also indicates that
the cation- interaction takes priority in crystal packing
in the mononuclear complex and no alternately layered
packing structure through arene-peruoroarene interactions
was obtained, in contrast to the case of H1.
In conclusion, we show three crystal structures of the
complexes 24, using the partially uorinated ligand 1.
The octahedral geometry of the Co
2+
and Ni
2+
ions gives
dinuclear complexes 2 and 3, as well as the corresponding
fully uorinated complexes 6 and 7, which are caused by
intramolecular stacking of the pentauorophenyl groups. On
the other hand, the square planar geometry of the Cu
2+
ion gives mononuclear complexes for all of the partially,
fully, and nonuorinated ligands (1, 5, and DBM) because
of the Jahn-Teller eect, which gives free cavity spaces
above the Cu
2+
ions. Thus, the phenyl group interacts to
the Cu
2+
ions through the cation- interaction for 4, as
well as [Cu(DBM)
2
]. The arene-peruoroarene and cation-
interactions are clearly shown as key electrostatic interactions
in the complexation behaviors and crystal structures of the
partially uorinated complexes.
Acknowledgment
This work was supported in part by a Kitasato University
Research Grant for Young Researchers.
References
[1] J.-M. Lehn, Supramolecular Chemistry, Wiley-VCH, Wein-
heim, Germany, 1995.
[2] G. R. Desiraju, Crystal Engineering: The Design of Organic
Solids, Elsevier, Amsterdam, The Netherlands, 1989.
[3] G. R. Desiraju, The Crystal as a Supramolecular Entity,
Perspective in Supramolecular Chemistry, vol. 2, John Wiley &
Sons, Chichester, UK, 1996.
[4] C. A. Hunter and J. K. M. Sanders, The nature of -
interactions, Journal of the American Chemical Society, vol.
112, no. 14, pp. 55255534, 1990.
[5] P. Hobza, H. L. Selzle, and E. W. Schlag, Structure and prop-
erties of benzene-containing molecular clusters: nonempirical
ab initio calculations and experiments, Chemical Reviews, vol.
94, no. 7, pp. 17671785, 1994.
[6] K. M uller-Dethlefs and P. Hobza, Noncovalent interactions:
a challenge for experiment and theory, Chemical Reviews, vol.
100, no. 1, pp. 143167, 2000.
[7] O. Yamauchi, A. Odani, and S. Hirota, Metal ion-assisted
weak interactions involving biological molecules. From small
complexes to metalloproteins, Bulletin of the Chemical Society
of Japan, vol. 74, no. 9, pp. 15251545, 2001.
[8] R. J. Doerksen and A. J. Thakkar, Quadrupole and octopole
moments of heteroaromatic rings, Journal of Physical Chem-
istry A, vol. 103, no. 48, pp. 1000910014, 1999.
[9] M. Nishio, Y. Umezawa, M. Hirota, and Y. Takeuchi, The
CH/ interaction: signicance in molecular recognition,
Tetrahedron, vol. 51, no. 32, pp. 86658701, 1995.
[10] J. C. Ma and D. A. Dougherty, The cation- interaction,
Chemical Reviews, vol. 97, no. 5, pp. 13031324, 1997.
[11] T. J. Shepodd, M. A. Petti, and D. A. Dougherty, Molecular
recognition in aqueous media: donor-acceptor and ion-dipole
interactions procedure tight binding for highly soluble guests,
Journal of the American Chemical Society, vol. 110, no. 6, pp.
19831985, 1988.
[12] R. A. Kumpf and D. A. Dougherty, A mechanism for ion
selectivity in potassium channels: computational studies of
cation- interactions, Science, vol. 261, no. 5129, pp. 1708
1710, 1993.
[13] C. R. Patrick and G. S. Prosser, A molecular complex of
benzene and hexauorobenzene, Nature, vol. 187, no. 4742,
p. 1021, 1960.
[14] J. H. Williams, The molecular electric quadrupole moment
and solid-state architecture, Accounts of Chemical Research,
vol. 26, no. 11, pp. 593598, 1993.
[15] A. F. M. Kilbinger and R. H. Grubbs, Arene-peruoroarene
interactions as physical cross-links for hydrogel formation,
Angewandte Chemie: International Edition, vol. 41, no. 9, pp.
15631566, 2002.
[16] A. S. Batsanov, J. C. Collings, and T. B. Marder, Arene-
peruoroarene interactions in crystal engineering. XV.
Ferrocene-decauorobiphenyl (1/1), Acta Crystallographica
C, vol. 62, no. 6, pp. m229m231, 2006.
[17] R. Xu, W. B. Schweizer, and H. Frauenrath, Soluble
poly(diacetylene)s using the pentauorophenyl-phenyl motif
as a supramolecular sunthon, Journal of the American Chem-
ical Society, vol. 130, no. 34, pp. 1143711445, 2008.
[18] M. Kim, T. J. Taylor, and F. P. Gabba, Hg(II) Pd(II)
metallophilic interactions, Journal of the American Chemical
Society, vol. 130, no. 20, pp. 63326333, 2008.
[19] D. Qui nonero, C. Garau, C. Rotger et al., Anion- inter-
actions: do they exist? Angewandte Chemie: International
Edition, vol. 41, no. 18, pp. 33893392, 2002.
8 International Journal of Inorganic Chemistry
[20] S. Demeshko, S. Dechert, and F. Meyer, Anion- interactions
in a carousel copper(II)-triazine complex, Journal of the
American Chemical Society, vol. 126, no. 14, pp. 45084509,
2004.
[21] P. De Hoog, P. Gamez, I. Mutikainen, U. Turpeinen, and J.
Reedijk, An aromatic anion receptor: anion- interactions do
exist, Angewandte Chemie: International Edition, vol. 43, no.
43, pp. 58155817, 2004.
[22] A. Hori, A. Shinohe, M. Yamasaki, E. Nishibori, S. Aoyagi,
and M. Sakata, 1 : 1 cross-assembly of two -diketonate
complexes through arene-peruoroarene interactions, Ange-
wandte Chemie: International Edition, vol. 46, no. 40, pp.
76177620, 2007.
[23] A. Hori and T. Arii, Cation- and arene-peruoroarene
interactions between Cu(II) uorine-substituted -diketonate
complex and benzenes, Crystal Engineering Community, vol.
9, no. 3, pp. 215217, 2007.
[24] A. Hori, A. Shinohe, S. Takatani, and T. K. Miyamoto,
Synthesis and crystal structures of uorinated -diketonate
metal (Al
3+
, Co
2+
, Ni
2+
, and Cu
2+
) complexes, Bulletin of the
Chemical Society of Japan, vol. 82, no. 1, pp. 9698, 2009.
[25] A. Hori and M. Mizutani, A benzene-rich pseudopoly-
morph of bis-[-1,3-bis-(penta-uoro-phen-yl) prop-ane-
1,3-dionato]-
3
O, O
,
: O
,
;
3
O : O, O
,
-bisaqua[1,3-bis(penta-
uorophenyl)propane-1,3-dionato-
2
O, O
,
]-nickel(II), Acta
Crystallographica C, vol. 65, no. 11, pp. m415m417, 2009.
[26] A. Hori, S. Takatani, T. K. Miyamoto, and M. Hasegawa,
Luminescence from - stacked bipyridines thorugh arene-
peruororene interactions, Crystal Engineering Community,
vol. 11, no. 4, pp. 567569, 2009.
[27] A. Hori and K. Naganuma, Fully and partially uorinated
avone derivatives, Acta Crystallographica, vol. C66, no. 5, pp.
o256o259, 2010.
[28] A. Hori, A. Akasaka, K. Biradha, S. Sakamoto, K. Yamaguchi,
and M. Fujita, Chirality induction through the reversible
catenation of coordination rings, Angewandte Chemie: Inter-
national Edition, vol. 41, no. 17, pp. 32693272, 2002.
[29] P. A. Vigato, V. Peruzzo, and S. Tamburini, The evolution of
-diketone or -diketophenol ligands and related complexes,
Coordination Chemistry Reviews, vol. 253, no. 7-8, pp. 1099
1201, 2009.
[30] D. V. Soldatov, A. T. Henegouwen, G. D. Enright, C. I.
Ratclie, and J. A. Ripmeester, Nickel(II) and zinc(II) diben-
zoylmethanates: molecular and crystal structure, polymor-
phism, and guest-or temperature-induced oligomerization,
Inorganic Chemistry, vol. 40, no. 7, pp. 16261636, 2001.
[31] L. David, C. Cr aciun, O. Cozar et al., Spectroscopic studies
of some oxygen-bonded copper(II) -diketonate complexes,
Journal of Molecular Structure, vol. 563-564, pp. 573578,
2001.
[32] B.-Q. Ma, S. Gao, Z.-M. Wang, C.-S. Liao, C.-H. Yan, and
G.-X. Xu, Synthesis and structure of bis(dibenzoylmethana-
to)copper(II), Journal of Chemical Crystallography, vol. 29,
no. 7, pp. 793796, 1999.
[33] F. A. Cotton and R. C. Elder, The crystal and molecular
structures of dimeric bis(acetylacetonato)aquocobalt(II) and
the preparation of some other new hydrates, Inorganic
Chemistry, vol. 5, no. 3, pp. 423429, 1966.
[34] C.-H. S. Wu, G. R. Rossman, H. B. Gray, G. S. Hammond, and
H. J. Schugar, Chelates of -diketones. VI. Synthesis and char-
acterization of dimeric dialkoxo-bridged iron(II) complexes
with acetylacetone and 2,2,6,6-tetramethylheptane-3,5-dione
(HDPM), Inorganic Chemistry, vol. 11, no. 5, pp. 990994,
1972.
[35] T. Choshi, S. horimoto, C. Y. Wang et al., Synthesis of
dibenzoylmethane derivatives and inhibition of mutagenicity
in Salmonella typhimurium, Chemical and Pharmaceutical
Bulletin, vol. 40, no. 4, pp. 10471049, 1992.
Hindawi Publishing Corporation
International Journal of Inorganic Chemistry
Volume 2011, Article ID 843051, 6 pages
doi:10.1155/2011/843051
Research Article
A Selective Chemosensor for Mercuric Ions Based on
4-Aminothiophenol-Ruthenium(II) Bis(bipyridine) Complex
Amer A. G. Al Abdel Hamid,
1
Mohammad Al-Khateeb,
2
Ziyad A. Tahat,
2
Mahmoud Qudah,
1
Safwan M. Obeidat,
1
and Abdel MonemRawashdeh
1
1
Department of Chemistry, Yarmouk University, Irbid 21163, Jordan
2
Department of Chemistry, JUST University, Irbid 22110, Jordan
Correspondence should be addressed to Amer A. G. Al Abdel Hamid, amerj@yu.edu.jo
Received 28 November 2010; Revised 22 February 2011; Accepted 23 February 2011
Academic Editor: W. T. Wong
Copyright 2011 Amer A. G. Al Abdel Hamid et al. This is an open access article distributed under the Creative Commons
Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is
properly cited.
A new ruthenium(II) complex (cis-ruthenium-bis[2,2

-bipyridine]-bis[4-aminothiophenol]-bis[hexauorophosphate]) has been


synthesized and characterized using standard analytical and spectroscopic techniques, FTIR,
1
H and
13
C-NMR, UV/vis, elemental
analysis, conductivity measurements, and potentiometric titration. Investigation of the synthesized complex with metal ions
showed that this complex has photochemical properties that are selective and sensitive toward the presence of mercuric ion in
aqueous solution. The detection limit for mercuric ions using UV/vis spectroscopy was estimated to be 0.4 ppm. The results
presented herein may have an important implication in the development of a spectroscopic selective detection for mercuric ions
in aqueous solution.
1. Introduction
Chemosensors that are selective for specic targets of metal
ions are continuously demanded. An especially important
category of chemosensors are those targeting toxic heavy
metal ions. Mercury in particular is a toxic metal, which
when accumulates in the vital organs of human and animals
causes poisonous eects that cause serious hematological
destruction, such as kidney malfunctioning and brain dam-
age [13]. Therefore, monitoring and precise determining
of mercuric ion concentration in water and thus in relevant
biological matrices are extremely benecial for the en-
vironmental and toxicological monitoring. Despite the
increasing eorts for developing low cost methodologies
for mercuric detection in aqueous solutions [416], the
tailored design of new mercuric-colorimetric chemosensors
that work eectively in aqueous media remains a key chal-
lenge. Up to now the great eort of researchers in this
eld is directed toward the development of new selective
chemosensors based on uorescence property [1733]. This
is of course due to many advantages in characterizing this
category of chemosensors, at top of which are the high selec-
tivity and sensitivity toward targeted metal ions. Although
these uorometric sensors have been employed enormously
in this eld, the colorimetric sensing [8, 28, 3436] of
metal ions has been shown to be less laboursome and
intensive alternative to uorescence techniques. The strong
thiophilicity of mercury is the most attractive property
that is usually taken into account when designing mercury
detection systems, whether they are based on colorimet-
ric or uorometric spectral changes [3540]. However,
investing this property in uorescent chemosensors is not
always wise and benecial, since many metals that are less
thiophilic than mercury (like silver, cadmium, and lead)
can promote reactions similar to those of mercury and
thus cause problematic sensing for mercury in relevant
environments. In addition, signicant challenges still exist
in this eld, especially those which are related to quench-
ing of uorescent signals by foreign interference presence.
Therefore, optimal ratiometric chemosensors (based on
colorimetric spectral changes) for mercuric ions that possess
fast response at ambient temperature, can selectively detect
2 International Journal of Inorganic Chemistry
N
N
N
N
Ru
N
SH
SH
N
H
H
H
H
(PF
6
)
2
Figure 1: Structure of the Cis-Ruthenium-bis[2,2

-bipyridine]-
bis[4-aminothiophenol]-bis[hexauorophosphate], (Ru-4-ASP).
3388
3476
1618
839
760
A
b
s
2650 2600 2550 2500
1335
35033427
Wavenumber (cm
1
)
4000 3000 2000 1000 400
(a)
(b)
Figure 2: IRspectra of (a) Ru(bpy)
2
Cl
2
2H
2
Oand (b) The complex
Ru-4-ASP. Enlargement of the SH bands in Ru-4-ASP in the
frequency range 25002650 cm
1
is shown in the inset.
Hg(II) ions, and can operate in aqueous media are still highly
demanded.
Polypyridine-Ru(II) complexes are interesting chro-
mophores. The assembly of oligopyridyl groups around
Ru(II) center yields mono- or polymetallic complex ions that
have been studied extensively for their potential ability as u-
orescent chemodetectors [4143] and energy or electron
transformers [4447]. Furthermore, these Ru(II)-metallo
complexes, in some cases, have demonstrated good photo-
physical and colorimetric sensing properties. Actually, this
feature has attracted our attention, and Ru(II)-bipyridine
derivative was employed in this study as a potential carrier
for preparation of new optical chemosensors.
In this paper, Ru(II)-bis(bipyridine) was selected to act as
the carrier for the sulfur-containing 4-aminothiophenol
moiety, which will act as the receptor for obtaining
a new chemosensor for mercuric ions. The incorporation
of the Hg(II) cations into thiols of the coordinated 4-
aminothiophenol ligand would be expected to inuence the
absorption properties of the Ru(II)-bipyridine core and thus
allowing access to a new potential chemosensor based on the
chromophore-spacer-receptor concept.
2. Experimental
2.1. General. All reagents were obtained from commercial
sources and were used without further purication. Column
250 nm
284 nm
413 nm
Wavelength (nm)
A
b
s
0.11
0.1
0.05
0
200 400 600 800
Figure 3: UV-vis absorption bands in Ru-4-ASP (1.0 M) complex.
chromatography was performed with Silica gel 60A/35
70 m, Merck Al
2
O
3
90 basic (0.0630.200 mm).
1
H and
13
C
NMR spectra were recorded on a Bruker 400 MHz. The
chemical shifts were reported using the residual solvent
signal as an indirect reference to TMS: acetone-d6 2.05 ppm
(
1
H) and 29.84 ppm (
13
C). UV/vis spectra were recorded for
50% ethanol/water (v/v) solutions on a Shimadzu UV-1800
spectrophotometer using 1 cm quartz cuvettes. IR measure-
ments were collected on a JASCO FT/IR-4100. All IR spectra
were recorded as pressed disks of the sample dispersed in
KBr powder. Typically, for each spectrum, 100 scans were
coadded at 4 cm
1
resolution. Microanalyses (C, H, N, S)
were performed using Euro EA elemental analyzer 3000.
Conductivity measurements were carried out using JENWAY
4010 conductivity meter employing 0.001 M solutions of
Ru-4-ASP complex. Potentiometric titration was carried
out using KHP standardized with 0.05 M NaOH solution
employing the Russell model RL150 Potentiometer. The pH
readings were taken after adding 1 mL of 0.05 M sodium
hydroxide increments allowing 30 seconds to pass to ensure
complete mixing before each pH measurement.
2.2. Synthesis
2.2.1. Cis-Ruthenium-bis[2,2

-bipyridine]-dichloride dihy-
drate,Ru(bpy)
2
Cl
2
2H
2
O]. It was prepared and character-
ized as reported, by Sullivan et al. [48]
2.2.2. Cis-Ruthenium-bis[2,2

-bipyridine]-bis[4-aminothio-
phenol]-bis[hexauorophosphate] (Ru-4-ASP). Under argon/
N
2
atmosphere, Ru(bpy)
2
Cl
2
2H
2
O (0.35 g, 0.67 mmol) and
4-aminothiophenol (4-ASP) (0.16 g, 1.4 mmol) were dis-
solved separately in a 1 : 1 mixture of absolute ethanol/water
(v/v). After mixing, the resulting solution was subjected to
reux overnight, after which NH
4
PF
6
(2.5 g, 13.5 mmol) dis-
solved in 2 mL water was added to precipitate the complex.
The solid product was collected by ltration and washed
throughly with 10 mL portions of water followed by diethyl
ether. The complex was then dried under reduced pressure to
yield 0.36 g (81%), m.p. 205207

C (dec.). Found C, 42.76;


H, 3.18; N, 9.31; S, 6.21 calc. for C
32
H
30
F
12
N
6
P
2
RuS
2
, C,
International Journal of Inorganic Chemistry 3
(1) 0.33 eq
3
2
5
4
7
6
1
Wavelength (nm)
A
b
s
(2) 0.66 eq
(3) 1 eq
(4) 1.33 eq
(5) 1.66 eq
(6) 2 eq
(7) 2.4 eq
0.3
0.2
0.1
0
200 220 240 260 280 300
Figure 4: The change in 250 nm absorption band of Ru-4-ASP
(1.0 M) as a function of the added 1.0 M Hg(II).
40.30; H, 3.17; N, 8.81; S, 6.72.
1
H-NMR 400 MHz (acetone-
d6) (ppm) 10.05(dd, J = 1.6, 5.6 Hz, 2H, H-6, bpy), 8.55
(dd, J = 1.2, 7.6 Hz, 2H, H-3, bpy), 8.47 (dd, J = 1.6, 6.7
Hz, 2H, H-3

, bpy), 8.05 (dd, J = 7.9, 8.6 Hz, 2H, H-4, bpy),


7.95 (dd, J = 7.9, 8.6 Hz, 2H, H-4

, bpy), 7.76 (dd, J = 1.6,


5.6 Hz, 2H, H-6

, bpy), 7.56 (dd, J = 5.0, 6.5 Hz, 2H, H-5,


bpy), 7.40 (dd, J = 5.0, 6.5 Hz, 2H, H-5

, bpy), 7.25 (dd, J =


1.3, 8.6 Hz, 2H, H-3

,5

), 6.80 (dd, J = 1.3, 9.0 Hz, 2H, H-


2

,6

).
13
C NMR (acetone-d6) (ppm) 158.9, 157.3, 156.6,
155.8, 152.4, 139.3, 139.2, 138.4, 138.0, 137.4, 132.3, 130.6,
128.8, 127.4, 127.3, 127.1, 126.3, 124.5, 124.3, 124.2, 123.9,
122.8; UVVis
max
() 250 nm (3.2 10
4
), 284 nm (3.7
10
4
), 413 nm (6.7 10
3
).
3. Results and Discussion
3.1. Synthesis of Ru-4-ASP Complex. Interaction of the
ruthenium complex, Ru(bpy)
2
Cl
2
2H
2
O, with 4-amino-
thiophenol (4-ASP) has resulted in Ru-4-ASP complex,
(Figure 1). This complex, Ru-4-ASP, appears to be ideal for
this application since we anticipated that Ru-4-ASP would
interact eciently with Hg(II) ions through the two thiols.
This interaction would facilitate spectral changes as a result
of mercury interaction. The Ru-4-ASPcomplex was prepared
in good yield by interacting Ru(bpy)
2
Cl
2
2H
2
O with 4-ASP
ligand, using one-step procedure. The formed complex was
dark brown-reddish colored and was observed to precipitate
as solid material immediately upon the addition of the
precipitating agent, ammonium hexauorophosphate.
The structure of Ru-4-ASP complex depicted in Figure 1
was veried by IR, H
1
, and
13
C-NMR, elemental analysis,
conductivity, and potentiometric titration. Introduction of
4-ASP ligand into Ru(bpy)
2
Cl
2
2H
2
O was evidenced by
the appearance of the two IR bands in the range 2560
2625 cm
1
. These two bands were assigned for the two
SH groups of the attached 4-ASP ligand. Since SH is
known as a weak IR absorber [49], the portion 2500
2650 cm
1
of the spectrum in Figure 2(b) has been blown
up to clearly show the two SH bands (see the inset in
Figure 2). The appearance of the NH-stretching vibrations
at 3388 and 3476 cm
1
in addition to the bending band at
Wavelength (nm)
A
b
s
Ru-4-ASP
413 nm Ru-4-ASP-Hg
II
505 nm
0.2
0.15
0.05
0
0.1
300 400 500 600
350 400 450 500
0.04
0.03
0.02
0.01
0
600 550
Figure 5: The shift in the 413 nm absorption band of Ru-4-ASP
(1.0 M) upon addition of 2 equivalents of Hg(II).
Wavelength (nm)
A
b
s
1.7
1.5
1
0
1.7
1.5
1
0.5
0
0.5
200 400 600 800
200 250 300 350 400 450
Ba
II
Ca
II
Mn
II
Co
II
Mg
II
Zn
II
Ni
II
Na
I
K
I
Figure 6: Absorption response of Ru-4-ASP (1.0 M) at 250 nm
upon the addition of Hg(II) co existed with 1.0 Mof Ba(II), Ca(II),
Co(II), Mg(II), Mn(II), Ni(II), Zn(II), Na(I), and K(I).
1618 cm
1
(Figure 2) indicates the coordination of the
0
1-
amino group of 4-ASP to the Ru(II) metal ion. The band at
1335 cm
1
, which was consistent with the C-N stretching in
0
1-amine, is another evidence for the coordination of 4-ASP
chelate to the Ru(bpy)
2
Cl
2
2H
2
O moiety via NH
2
group.
In 50% water/ethanol (v/v) solution, Ru-4-ASP turned to
light brown-reddish-colored solution and gave rise to three
absorption bands, two strong bands in the UV-region, at 250
and 284 nm, and a third weak band in the visible region at
413 nm; these bands are shown in Figure 3.
1
H- and
13
C-NMR, in terms of the peak shift and number
of carbon atoms along with the elemental analysis of carbon,
hydrogen, nitrogen, and sulfur are consistent with the
structure of Ru-4-ASP proposed in Figure 1. On the other
hand, the potentiometric titration of one equivalent of Ru-
4-ASP with 0.05 M NaOH indicated that thiol groups lost
their protons completely in the pH range of 6.0 to 8.8. This
range corresponds to two equivalents of sodium hydroxide.
Furthermore, the deprotonation intervals of both SH groups
were observed to be indistinguishable, where no well-dened
sharp equivalence points were noticed; instead, the two
regions were overlapped due to the high matching and sim-
ilarity between the two groups in the Ru-4-ASP structure.
4 International Journal of Inorganic Chemistry
3.2. Interaction of Hg(II) with Ru-4-ASP Dye. A spectro-
scopic titration of Hg(II) was conducted with 1.0 M
solution of Ru-4-ASP in 50% ethanol/water (v/v) solution
at pH 6. Upon addition of an increased amount of Hg(II)
ions, the area under the UV-absorption peak corresponding
to 250 nm starts to decline, while the band at 284 nm
stayed in place without any noticeable change, Figure 4. The
reaction responsible for these changes was observed to reach
completion within <30 seconds and the observed reduction
in the 250 nm absorption band was proportional to the
added Hg(II) concentration. Moreover, the saturation of
Ru-4-ASP with Hg(II) ions was attained after adding 2.0
equivalents of Hg(II); beyond this point, more additions
of Hg(II) brought no spectral changes in the absorption
proles as shown in Figure 4. On the other hand, the visible
band at 413 nm was observed to undergo a red shift (into
higher wavelength) upon adding two equivalents of Hg(II),
Figure 5. At this point, the color of the Ru-4-ASP solution
treated with Hg(II) was visually observed, at once, to change
from light brown-reddish to orange-reddish. This change in
color was not noticed to take place in solution unless the
amount of the added Hg(II) reaches the equivalent amount.
In a control experiment, Ru-4-ASP was found to retain its
original color when treated with only 1.0 equivalent of
Hg(II), and this color was found to be stable over time as was
observed when the solution was kept on the shelf for more
than three weeks.
Similar spectral changes were observed when other salts
of mercury, such as mercuric nitrate and mercuric perchlo-
rate, were used. Therefore, it appears that counter anions
accompanying Hg(II) ion have negligible eect on the sen-
sation activity of Ru-4-ASP complex. Moreover, the color
change in the presence of Hg(II) was found to be insensitive
to interferences by other metal cations. That was observed
when the Ru-4-ASP complex was allowed to interact with
solutions containing submicromolar amounts of Hg(II) and
micromolar quantities of the metal cations: Ag(I), Zn(II),
Cu(II), Pb(II), Cd(II), Ni(II), Co(II), Fe(II), Fe(III), Mn(II),
Mg(II), Ca(II), Ba(II), Li(I), K(I), Na(I), Rh(III), Cr(II),
Cr(III). The absorption proles of some of these ions are
shown in Figure 6.
On the other hand, when the optical measurements were
repeated for the above cationic solutions but in absence of
Hg(II), no change was observed in the three absorption peaks
of Ru-4-ASP, Figure 7. However, higher absorbance values
were noticed for the 250 nm band when the Ru-4-ASP
complex was interacted with Fe(III), Rh(III), and Cr(III).
Compared to the divalent ions, the high absorbance values
observed with these ions are attributed to the high electron
deciency of the three trivalent ions and consequently the
stronger interaction they exhibit with the electron-rich lig-
and. Fortunately, in all cases, no spectral changes (peak
reduction) similar to those witnessed when Ru-4-ASP was
interacted with Hg(II) have been induced, Figure 8.
Therefore, the outcomes of the proceeded optical mea-
surements of Ru-4-ASP with Hg(II) demonstrate the lack of
interferences by other metal cations or their accompanied
anions. This means that Ru-4-ASP complex has a remarkable
selectivity toward Hg(II) ion over other metal ions. To see
(a)
(a)
(b)
(b)
0.3
0.2
0.1
0
0.3
0.2
0.1
0
200 250 300 350 400 450
Wavelength (nm)
200 400 600 800
A
b
s
Ba
II
Ca
II
Mn
II
Co
II
Mg
II
Zn
II
Ni
II
Na
I
K
I
Figure 7: Absorption spectra of Ru-4-ASP (1.0 M) at 250 nm, (a)
in the presence of 1.0 M of Ba(II), Ca(II), Co(II), Mg(II), Mn(II),
Ni(II), Zn(II), Na(I), and K(I) and (b) Ru-4-ASP (1.0 M) with no
additions.
Cr
III
Rh
III
Fe
III
Divalent cations
A
b
s
0.5
0.4
0.3
0.2
0.1
0
0.5
0.4
0.3
0.2
0.1
0
200 250 300 350 400 450
Wavelength (nm)
200 400 600 800
Figure 8: The change in absorption response of Ru-4-ASP (1.0 M)
at 250 nm upon the addition of 1.0 M of trivalent ions: Cr(III),
Fe(III), and Rh(III) compared to the divalent ions: Ba(II), Ca(II),
Co(II), Mg(II), Mn(II), Ni(II), Zn(II), Na(I), and K(I).
P
e
a
k
a
r
e
a
0.1
0.08
0.06
0.04
0.02
0
0.02
0.04
0.06
0.08
0.1
0 0.3 0.6 0.9 1.2 1.5 1.8 2.1 2.4 2.7
Equivalent Hg(II)
Figure 9: Titration of the change in the area of the 250 nm
absorption peak of Ru-4-ASP versus equivalents of Hg(II).
International Journal of Inorganic Chemistry 5
the practical applicability of this system as a detector for
mercuric ions, the detection limit was evaluated. The titra-
tion prole of Ru-4-ASP with Hg(II), which is shown in
Figure 9, demonstrates that the detection of Hg(II) is below
part per-million level (<0.4 ppm).
In summary, this investigation described in the pro-
ceeding has resulted in a development of a highly selective
and sensitive Ru-4-ASP-based chemosensor for mercuric ion
detection in aqueous medium. This system works eciently
with remarkable high selectivity and sensitivity and under
conditions similar to those encountered in relevant Hg(II)-
contaminated environments. In such environments, Hg(II)
coexists in a matrix of other interfering ions. The ndings
of this investigation suggest that this strategy would serve
as a foundation for practical, rapid detection and precise
determination of Hg(II) ions in aqueous environments.
4. Conclusion
The results presented in this investigation demonstrate that,
the interaction taking place between Hg(II) ions and Ru-
4-ASP complex (through thiol groups) was responsible for
the observed spectral changes of the corresponding Ru-4-
ASP dye. Interestingly, the investigation described above has
resulted in a development of a highly selective and sensitive
chemosensor for Hg(II) ions in alcoholicaqueous solution
even in the presence of relatively high concentrations of
potentially competing other metal cations. This includes
those cations that are identied by the U.S. Environmental
Protection Agency as potential environmental water pol-
lutants [50] such as Zn(II), Cd(II), Pb(II), Ni(II), and
Fe(II). Furthermore, and in terms of sensitivity, the limit of
quantication of the system, based on UV-vis spectroscopy
measurements, was estimated to be lower than 0.4 ppm,
providing a good chemosensation for Hg(II) ions.
Acknowledgments
This work was nancially supported by the Faculty of Grad-
uate Studies and Scientic Research Yarmouk University.
The authors express great thanks for the Department of
Chemistry-Jordan University of Science and Technology
(JUST) for providing instrumentation.
References
[1] F. Monnet-Tschudi, M. G. Zurich, C. Boschat, A. Corbaz, and
P. Honegger, Involvement of environmental mercury and
lead in the etiology of neurodegenerative diseases, Reviews on
Environmental Health, vol. 21, no. 2, pp. 105117, 2006.
[2] J. Mutter and J. Naumann, Blood mercury levels and neu-
robehavior, Journal of the American Medical Association, vol.
294, no. 6, p. 679, 2005.
[3] J. Watts, Mercury poisoning victims could increase by
20,000, The Lancet, vol. 358, no. 9290, p. 1349, 2001.
[4] E. Palomares, R. Vilar, and J. R. Durrant, Heterogeneous col-
orimetric sensor for mercuric salts, Chemical Communica-
tions, vol. 10, no. 4, pp. 362363, 2004.
[5] L. F. Capit an-Vallvey, C. Cano Raya, E. L opez L opez, and M. D.
Fern andez Ramos, Irreversible optical test strip for mercury
determination based on neutral ionophore, Analytica Chim-
ica Acta, vol. 524, no. 1-2, pp. 365372, 2004.
[6] S. S. M. Hassan, W. H. Mahmoud, A. H. K. Mohamed, and
A. E. Kelany, Mercury(II) ion-selective polymeric membrane
sensors for analysis of mercury in hazardous wastes, Analyti-
cal Sciences, vol. 22, no. 6, pp. 877881, 2006.
[7] C. Cano-Raya, M. D. Fern andez-Ramos, J. G omez-S anchez,
and L. F. Capit an-Vallvey, Irreversible optical sensor for mer-
cury determination based on tetraarylborate decomposition,
Sensors and Actuators, vol. 117, no. 1, pp. 135142, 2006.
[8] E. Coronado, J. R. Gal an-Mascar os, C. Mart-Gastaldo et al.,
Reversible colorimetric probes for mercury sensing, Journal
of the American Chemical Society, vol. 127, no. 35, pp. 12351
12356, 2005.
[9] J. Wang, X. Qian, and J. Cui, Detecting Hg
2+
ions with an ICT
uorescent sensor molecule: remarkable emissionspectra shift
and unique selectivity, Journal of Organic Chemistry, vol. 71,
no. 11, pp. 43084311, 2006.
[10] Y. Tang, F. He, M. Yu et al., A reversible and highly selective
uorescent sensor for mercury(II) using poly(thiophene)s that
contain thymine moieties, Macromolecular Rapid Communi-
cations, vol. 27, no. 6, pp. 389392, 2006.
[11] S. Ou, Z. Lin, C. Duan, H. Zhang, and Z. Bai, A sugar-
quinoline uorescent chemosensor for selective detection of
Hg
2+
ion in natural water, Chemical Communications, no. 42,
pp. 43924394, 2006.
[12] Y. Yu, L.-R. Lin, K.-B. Yang, X. Zhong, R.-B. Huang, and L.-S.
Zheng, p-Dimethylaminobenzaldehyde thiosemicarbazone:
a simple novel selective and sensitive uorescent sensor for
mercury(II) in aqueous solution, Talanta, vol. 69, no. 1, pp.
103106, 2006.
[13] S. M. Cheung and W. H. Chan, Hg
2+
sensing in aqueous solu-
tions: an intramolecular charge transfer emission quenching
uorescent chemosensors, Tetrahedron, vol. 62, no. 35, pp.
83798383, 2006.
[14] J. Wang and X. Qian, A series of polyamide receptor based
PET uorescent sensor molecules: positively cooperative Hg
2+
ion binding with high sensitivity, Organic Letters, vol. 8, no.
17, pp. 37213724, 2006.
[15] E. M. Nolan, M. E. Racine, and S. J. Lippard, Selective
Hg(II) detection in aqueous solution with thiol derivatized
uoresceins, Inorganic Chemistry, vol. 45, no. 6, pp. 2742
2749, 2006.
[16] Y. Zhao and Z. Zhong, Detection of Hg
2+
in aqueous solu-
tions with a foldamer-based uorescent sensor modulated by
surfactant micelles, Organic Letters, vol. 8, no. 21, pp. 4715
4717, 2006.
[17] E. M. Nolan and S. J. Lippard, A turn-on uorescent sensor
for the selective detection of mercuric ion in aqueous media,
Journal of the American Chemical Society, vol. 125, no. 47, pp.
1427014271, 2003.
[18] A. Ono and H. Togashi, Highly selective oligonucleotide-
based sensor for mercury(II) in aqueous solutions, Ange-
wandte ChemieInternational Edition, vol. 43, no. 33, pp.
43004302, 2004.
[19] J. V. Mello and N. S. Finney, Reversing the discovery par-
adigm: a new approach to the combinatorial discovery of u-
orescent chemosensors, Journal of the American Chemical
Society, vol. 127, no. 29, pp. 1012410125, 2005.
[20] B. Valeur, Molecular Fluorescence: Principles and Applications,
Wiley-VCH, Weinheim, Germany, 2002.
[21] A. P. de Silva, H. Q. Gunaratne, T. Gunnlaugsson et al.,
Signaling recognition events with uorescent sensors and
6 International Journal of Inorganic Chemistry
switches, Chemical Reviews, vol. 97, no. 5, pp. 15151566,
1997.
[22] L. Fabbrizzi and A. Poggi, Sensors and switches from
supramolecular chemistry, Chemical Society Reviews, vol. 24,
no. 3, pp. 197202, 1995.
[23] A. W. Czarnik, Chemical communication in water using
uorescent chemosensors, Accounts of Chemical Research, vol.
27, no. 10, pp. 302308, 1994.
[24] S.-Y. Moon, N. J. Youn, S. M. Park, and S.-K. Chang, Diamet-
rically disubstituted cyclam derivative having Hg
2+
-selective
uoroionophoric behaviors, Journal of Organic Chemistry,
vol. 70, no. 6, pp. 23942397, 2005.
[25] Q.-Y. Chen and C.-F. Chen, A new Hg
2+
-selective uorescent
sensor based on a dansyl amide-armed calix[4]-aza-crown,
Tetrahedron Letters, vol. 46, no. 1, pp. 165168, 2005.
[26] X. G. Guo, X. Qian, and L. Jia, A highly selective and sensitive
uorescent chemosensor for Hg
2+
in neutral buer aqueous
solution, Journal of the American Chemical Society, vol. 126,
no. 8, pp. 22722273, 2004.
[27] J. Y. Kwon, J. H. Soh, Y. J. Yoon, and J. Yoon, Highly eective
uorescent sensor for Hg
2+
in aqueous solution, Supramolec-
ular Chemistry, vol. 16, no. 8, pp. 621624, 2004.
[28] J. V. Ros-Lis, R. Martnez-M a nez, K. Rurack, F. Sancen on, J.
Soto, and M. Spieles, Highly selective chromogenic signaling
of Hg
2+
in aqueous media at nanomolar levels employing a
squaraine-based reporter, Inorganic Chemistry, vol. 43, no. 17,
pp. 51835185, 2004.
[29] A. B. Descalzo, R. Martnez-M a nez, R. Radeglia, K. Rurack,
and J. Soto, Coupling selectivity with sensitivity in an inte-
grated chemosensor framework: design of a Hg
2+
-responsive
probe, operating above 500 nm, Journal of the American
Chemical Society, vol. 125, no. 12, pp. 34183419, 2003.
[30] M. J. Choi, M. Y. Kim, and S. K. Chang, A new Hg
2+
-selective
chromoionophore based on calix[4]arenediazacrown ether,
Chemical Communications, no. 17, pp. 16641665, 2001.
[31] L. Prodi, C. Bargossi, M. Montalti et al., An eective uores-
cent chemosensor for mercury ions, Journal of the American
Chemical Society, vol. 122, no. 28, pp. 67696770, 2000.
[32] K. Rurack, M. Kollmannsberger, U. Resch-Genger, and J.
Daub, A selective and sensitive uoroionophore for Hg(II),
Ag(I), and Cu(II) with virtually decoupled uorophore and
receptor units, Journal of the American Chemical Society, vol.
122, no. 5, pp. 968969, 2000.
[33] H. Sakamoto, J. Ishikawa, S. Nakao, and H. Wada, Excel-
lent mercury(II) ion selective uoroionophore based on a
3,6,12,15-tetrathia-9-azaheptadecane derivative bearing a ni-
trobenzoxadiazolyl moiety, Chemical Communications, no.
23, pp. 23952396, 2000.
[34] J. H. Huang, W. H. Wen, Y. Y. Sun, P. T. Chou, and J. M. Fang,
Two-stage sensing property via a conjugated donor-acceptor-
donor constitution: application to the visual detection of
mercuric ion, Journal of Organic Chemistry, vol. 70, no. 15,
pp. 58275832, 2005.
[35] M. Y. Chae and A. W. Czarnik, Fluorometric chemodosime-
try. Mercury(II) and silver(I) indication in water via enhanced
uorescence signaling, Journal of the American Chemical
Society, vol. 114, no. 24, pp. 97049705, 1992.
[36] V. Dujols, F. Ford, and A. W. Czarnik, A long-wavelength
uorescent chemodosimeter selective for Cu(II) ion in water,
Journal of the American Chemical Society, vol. 119, no. 31, pp.
73867387, 1997.
[37] J. V. Ros-Lis, M. D. Marcos, R. M artinez-M a nez, K. Rurack,
and J. Soto, A regenerative chemodosimeter based on metal-
induced dye formation for the highly selective and sensitive
optical determination of Hg
2+
ions, Angewandte Chemie
International Edition, vol. 44, no. 28, pp. 44054407, 2005.
[38] G. Hennrich, H. Sonnenschein, and U. Resch-Genger, Redox
switchable uorescent probe selective for either Hg(II) or
Cd(II) and Zn(II), Journal of the American Chemical Society,
vol. 121, no. 21, pp. 50735074, 1999.
[39] G. Zhang, D. Zhang, S. Yin, X. Yang, Z. Shuai, and D. Zhu,
1,3-Dithiole-2-thione derivatives featuring an anthracene
unit: new selective chemodosimeters for Hg(II) ion, Chemical
Communications, no. 16, pp. 21612163, 2005.
[40] B. Liu and H. Tian, Aselective uorescent ratiometric chemo-
dosimeter for mercury ion, Chemical Communications, no.
25, pp. 31563158, 2005.
[41] C.-Y. Li, X.-B. Zhang, Z. Jin, R. Han, G.-L. Shen, and R.-Q. Yu,
A uorescent chemosensor for cobalt ions based on a multi-
substituted phenol-ruthenium(II) tris(bipyridine) complex,
Analytica Chimica Acta, vol. 580, no. 2, pp. 143148, 2006.
[42] M. E. Padilla-Tosta, J. M. Lloris, R. Martnez-M a nez et al.,
Bis(terpyridyl)-ruthenium(II) units attached to polyazacy-
cloalkanes as sensing uorescent receptors for transition metal
ions, European Journal of Inorganic Chemistry, no. 4, pp. 741
748, 2000.
[43] M. D. Pratt and P. D. Beer, Heterodinuclear ruthenium(II)
bipyridyl-transition metal dithiocarbamate macrocycles for
anion recognition and sensing, Tetrahedron, vol. 60, no. 49,
pp. 1122711238, 2004.
[44] L. Sun, L. Hammarstr om, B.

Akermark, and S. Styring,
Towards articial photosynthesis: rutheniummanganese
chemistry for energy production, Chemical Society Reviews,
vol. 30, pp. 3649, 2001.
[45] M. Sj odin, S. Styring, H. Wolpher, Y. Xu, L. Sun, and L.
Hammarstr om, Switching the redox mechanism: models for
proton-coupled electron transfer from tyrosine and trypto-
phan, Journal of the American Chemical Society, vol. 127, no.
11, pp. 38553863, 2005.
[46] V. Balzani, A. Juris, and M. Venturi, Luminescent and redox-
active polynuclear transition metal complexes, Chemical
Reviews, vol. 96, no. 2, pp. 759833, 1996.
[47] C. Kaes, A. Katz, and M. W. Hosseini, Bipyridine: the most
widely used ligand. A review of molecules comprising at least
two 2,2-bipyridine units, Chemical Reviews, vol. 100, no. 10,
pp. 35533590, 2000.
[48] B. P. Sullivan, D. J. Salmon, and T. J. Meyer, Mixed phosphine
2,2-bipyridine complexes of ruthenium, Inorganic Chem-
istry, vol. 17, no. 12, pp. 33343341, 1978.
[49] K. N. Solomon, Infrared Absorption Spectroscopy, Holden-Day,
2nd edition, 1977.
[50] Updates on water contaminants, www.epa.gov/safewater/hfacts
.html.

Você também pode gostar