ArticlePDF Available

Cobalt(III) Complexes with Thiosemicarbazones as Potential anti-Mycobacterium tuberculosisAgents

Authors:

Abstract and Figures

CoIII complexes derived from 2-acetylpyridine N(4)-R thiosemicarbazone (Hatc-R, R = alkyl, aryl) have been characterized by elemental analysis, FTIR, UV-Visible and 1 H NMR spectroscopies, cyclic voltammetry (CV), conductimetry measurements and single crystal X-ray diffractometry. The results obtained are consistent with the oxidation of the CoII center to CoIII upon coordination of the monoanionic N,N,S-tridentate thiosemicarbazone ligands, resulting in octahedral ionic complexes of the type [Co(atc-R)2]Cl. Electrochemistry studies show two reversible processes referring to the redox couples CoIII/CoII and CoII/CoI which can be modified by the inductive effects of the substituents groups at the N4 position of the ligands. Two CoIII complexes showed satisfactory activity with minimal inhibitory concentration value under 10 μmol L–1 and one presented quite low cytotoxicity against VERO and J774A.1 cells (IC50), resulting in high selectivity index (SI > 10).
Content may be subject to copyright.
Article
J. Braz. Chem. Soc., Vol. 25, No. 10, 1848-1856, 2014.
Printed in Brazil - ©2014 Sociedade Brasileira de Química
0103 - 5053 $6.00+0.00
Ahttp://dx.doi.org/10.5935/0103-5053.20140149
*e-mail: deflon@iqsc.usp.br
Cobalt(III) Complexes with Thiosemicarbazones as Potential
anti‑Mycobacterium tuberculosis Agents
Carolina G. Oliveira,a Pedro Ivo da S. Maia,b Marcelo Miyata,c Fernando R. Pavan,c
Clarice Q. F. Leite,c Eduardo Tonon de Almeidad and Victor M. Deflon*,a
aInstituto de Química de São Carlos, Universidade de São Paulo, 13566-590 São Carlos-SP, Brazil
bDepartamento de Química, Universidade Federal do Triângulo Mineiro, 38025-440 Uberaba-MG, Brazil
cFaculdade de Ciências Farmacêuticas, Universidade Estadual Paulista, 14801-902 Araraquara-SP, Brazil
dInstituto de Química, Universidade Federal de Alfenas, 37130-000 Alfenas-MG, Brazil
Complexos de CoIII derivados da 2-acetilpiridina N(4)-R tiossemicarbazona (Hatc-R, R = alquil
ou aril) foram caracterizados por análise elementar, espectroscopia na região do infravermelho,
UV-Visível e 1H RMN, voltametria cíclica (VC), medidas de condutividade e difração de raios X
em monocristal. Os resultados obtidos são consistentes com a oxidação do centro de CoII para CoIII
após a coordenação N,N,S-tridentada e monoaniônica dos ligantes tiossemicarbazonas, resultando
em complexos octaédricos iônicos do tipo [Co(atc-R)2]Cl. Os estudos de eletroquímica mostram
dois processos reversíveis, referentes aos pares redox CoIII/CoII e CoII/CoI, que são afetados pelo
efeito indutivo dos grupos substituintes na posição N4 dos ligantes. Dois complexos de CoIII se
mostraram satisfatoriamente ativos, com valores de concentração inibitória mínima abaixo de
10 μmol L–1 e um deles apresentou muito baixa citotoxicidade contra células VERO e J774A.1
(IC50), conferindo-lhe altos índices de seletividade (SI > 10).
CoIII complexes derived from 2-acetylpyridine N(4)-R thiosemicarbazone (Hatc-R, R = alkyl,
aryl) have been characterized by elemental analysis, FTIR, UV-Visible and 1H NMR spectroscopies,
cyclic voltammetry (CV), conductimetry measurements and single crystal X-ray diffractometry. The
results obtained are consistent with the oxidation of the CoII center to CoIII upon coordination of the
monoanionic N,N,S-tridentate thiosemicarbazone ligands, resulting in octahedral ionic complexes
of the type [Co(atc-R)2]Cl. Electrochemistry studies show two reversible processes referring to
the redox couples CoIII/CoII and CoII/CoI which can be modified by the inductive effects of the
substituents groups at the N4 position of the ligands. Two CoIII complexes showed satisfactory
activity with minimal inhibitory concentration value under 10 μmol L–1 and one presented quite low
cytotoxicity against VERO and J774A.1 cells (IC50), resulting in high selectivity index (SI > 10).
Keywords: cobalt(III), thiosemicarbazones, anti-mycobacterium tuberculosis activity
Introduction
Tuberculosis (TB) still causes the death of millions
of people every year, more than any other disease around
the world, as described by the World Health Organization
(WHO).1 The pathogenicity of Mycobacterium tuberculosis
(MTB), the human pathogen responsible for TB, is
based on the development of methods to survive inside
host cells, comprising the capacity to take possession
of macrophages.2 Despite a 95% efficacious 6-month
treatment, the TB problem is still expanding world-wide.3
It’s estimated that one-third of the world’s population
is infected with dormant forms of MTB, 10% of which
will develop the disease among their lives.4 The actual
research on TB has been focused on the increased number
of multidrug and extensively drug-resistant TB (MDR- and
XDR-TB),5 especially in HIV-positive patients, giving
rise to very high mortality.6 The number of promising
anti-TB drugs following pre-clinical tests has increased
and although they involve diverse possible mechanisms, no
one targets dormant bacteria, which means that the latent
infections cannot be eliminated by the moment.7 Therefore,
Oliveira et al. 1849Vol. 25, No. 10, 2014
it is mandatory to develop new anti-MTB agents that can
solve the current therapy problems and inhibit the growth
of pathogenic microorganisms in their latent forms.
Biologically active transition metal complexes
constitute an increasing research area.8-11 Metals like Mn,
Fe, Co, Ni, Cu and Zn, even in a very low concentration
in human body, are responsible for a series essential of
biological functions10 being studied in the bioinorganic
chemistry field.12-14 Cobalt is an essential element for life
being the main component of Vitamin B12, which is an
essential micronutrient that is required for human health
and, more importantly, is required in large quantities by
cells that are replicating DNA prior to cell division.15
Moreover, a lot of new cobalt compounds have been
demonstrated to possess particular biological activities,
such as antitumor and antibacterial.16-20
The interest on the thiosemicarbazones (TSCs) and their
metal complexes is due to their versatile chemistry21 and
pharmacological activities,22-26 which also include anti-TB
properties.11 Frequently, the biological activity presented
by the free thiosemicarbazone ligands is enhanced upon
complexation.27 In an effort to select lead candidates for
treatment of TB, in the last few years we have studied
and reported complexes derived from thiosemicarbazones
which presented a high anti-MTB activity.11 Precisely,
VIV and VV,8 NiII 14 and MnII complexes28 derived from
2-acetylpyridine thiosemicarbazones have shown excellent
activity against MTB. In this context, we focused our
interest in a continuation of the previous studies with some
of the first transition metal row by developing now new
CoIII compounds, cationic instead of neutral complexes
as the NiII and MnII compounds studied before.14,28 Hence,
believing in the high potential of such compounds as
anti-MTB agents and the biochemical features of cobalt,
here we describe the preparation of CoIII complexes derived
from 2-acetylpyridine-thiosemicarbazones, their full
characterization as well as the study of their anti-MTB
activity and cytotoxicity.
Experimental
Materials
2-Acetylpyridine, thiosemicarbazide, 4-methyl-
3-thiosemicarbazide, 4-ethyl-3-thiosemicarbazide,
4-phenyl-3-thiosemicarbazide and analytical reagents
grade chemicals and solvents were obtained commercially
and used without further purification. 4-Cyclohexyl-3-
thiosemicarbazide was prepared as previously described.27
The ligands Hatc, Hatc-Me, Hatc-Et, Hatc-Ch and Hatc-Ph
were prepared by refluxing equimolar ethanolic solutions
containing the desired thiosemicarbazide (10 mmol)
and 2-acetylpyridine (10 mmol) for 1 h, as reported
elsewhere.29,30
Instruments
FTIR spectra were measured as KBr pellets on a
Shimadzu IR Prestige-21 spectrophotometer between
400 and 4000 cm−1. Elemental analyses were determined
using a Leco Instrument, model Truspec CHNS-O.
The conductivities of the complexes were measured in
1 × 10-3 mol L–1 MeOH or H2O solutions using an Orion Star
Series conductometer. UV-visible (UV-Vis) spectra were
measured with a Shimadzu UV-1800 spectrophotometer
in MeOH solutions. The electrochemical experiments
were carried out at room temperature in acetonitrile
containing 0.1 mol L–1 tetrabutylammonium perchlorate
(PTBA) (Fluka Purum) as supporting electrolyte, using an
electrochemical analyzer μAutolab III. Cyclic voltammetry
experiments were performed with a glassy carbon (CG)
working stationary electrode, a platinum auxiliary electrode
and an aqueous Ag/AgCl reference, carried out with a rate
sweep of 100 mV s–1 or 200 mV s–1. The 1H NMR spectra
were acquired using equipment Agilent 400/54 Premium
Shielded 9.4 T, working at 399.8 MHz for 1H. The NMR
spectra were internally referenced to TMS.
Crystal structure determination
Brown crystals of [Co(atc)2]Cl∙H2O (1∙H2O) were
grown by slow evaporation of the ethanol mother solution
of 1. Crystals of [Co(atc-Ph)2]∙MeOH (5∙MeOH) were
obtained by recrystallization of 5 in MeOH/CH2Cl2 1:2 at
room temperature. The data collections were performed
using Mo-Kα radiation (λ = 71.073 pm) on a BRUKER
APEX II Duo diffractometer. Standard procedures were
applied for data reduction and absorption correction.
The structures were solved with SHELXS97 using direct
methods,31 and all non-hydrogen atoms were refined with
anisotropic displacement parameters with SHELXL97.32
The hydrogen atoms were calculated at idealized positions
using the riding model option of SHELXL97.32 Table 1
presents more detailed information about the structures
determination.
Determination of MICs
The anti-MTB activity of the compounds was
determined as MIC (minimal inhibitory concentration)
values by the REMA (Resazurin Microtiter Assay) method
according to Palomino et al.33
Cobalt(III) Complexes with Thiosemicarbazones as Potential anti-Mycobacterium tuberculosis Agents J. Braz. Chem. Soc.
1850
Cytotoxicity assay
In vitro cytotoxicity assays (IC50, half maximal
inhibitory concentration) were performed first on VERO
epithelial cells (ATCC CCL81). Following this approach,
the most selective compound (higher SI) was additionally
tested on the J774A.1 (ATCC TIB-67) murine macrophage
cell line. Both studies were recorded as reported before in
a previously work.28
Selectivity index
The selectivity index (SI) was calculated by dividing
IC50 for VERO cells by the MIC for the pathogen; if
SI 10, the compound is considered suitable for further
investigations.28
Preparations
The CoIII complexes were synthesized by adding
0.25 mmol CoCl2∙6H2O to solutions of 0.5 mmol of the
desired ligands in EtOH (15 mL). The resulting solutions
were stirred for 2 h under reflux. The solutions were kept
under slow evaporation at room temperature until brown
precipitates were formed. After 3 days the solids were
filtered off, washed with hexane and dried under vacuum.
[Co(atc)2]Cl·H2O (1·H2O)
Yield 0.085 g (68%). Analysis: Calc. for
C16H20ClCoN8OS2: C, 38.52; H, 4.04; N, 22.46%. Found:
C, 39.58; H, 4.19; N, 22.87%. IR (KBr) ν/cm−1 3259, 3101,
1620, 1598, 1035, 773; UV-Vis in 3.2 × 10–5 mol L–1 MeOH
solution [λ / nm (ε / L mol–1 cm–1)]: 424 (4801), 365 (7289),
Table 1. Crystal data and structure refinement for [Co(atc)2]Cl∙H2O (1∙H2O) and [Co(atc-Ph)2]Cl∙MeOH (5∙MeOH)
1∙H2O5∙MeOH
Empirical formula C16H20ClCoN8OS2C29H30ClCoN8OS2
Formula weight 498.90 665.11
Temperature / K 296(2) 293(2)
Wavelength / Å 0.71073 0.71073
Crystal system Monoclinic Monoclinic
Space group P21/n P21/c
Unit cell dimensions a = 10.0774(4) / Å α = 90°
b = 17.4487(7) / Å β = 91.012(2)°
c = 11.6658(4) / Å γ = 90°
a = 9.9167(2) / Å α = 90°
b = 23.4163(5) / Å β = 121.17(10)°
c = 15.2784(3) / Å γ = 90°
Volume / Å32050.97(14) 3035.55(11)
Z 4 4
Density (calculated) / (mg m–3) 1.616 1.455
Absorption coefficient / mm–1 1.197 0.829
F(000) 1024 1376
Crystal size / mm30.19 × 0.17 × 0.06 0.971 × 0.386 × 0.37
Theta range for data collection 2.10 to 25.35° 1.74 to 25.05°
Index ranges −12 h 11,
−20 k 20,
−14 l 14
−11 h 11,
−27 k 24,
−16 l 18
Reflections collected 12598 18887
Independent reflections [R(int) = 0.0364] [R(int) = 0.0181]
Completeness to theta = 25.14° / % 99.3 99.2
Absorption correction Semi-empirical from equivalents Semi-empirical from equivalents
Max. and min. transmission 0.7452 and 0.6388 0.7452 and 0.6775
Refinement method Full-matrix least-squares on F2Full-matrix least-squares on F2
Data / restraints / parameters 3725 / 3 / 270 5327 / 0 / 382
Goodness-of-fit on F21.037 1.068
Final R indices [I>2sigma(I)] R1 = 0.0402, wR2 = 0.0981 R1 = 0.0316, wR2 = 0.0864
R indices (all data) R1 = 0.0642, wR2 = 0.1112 R1 = 0.0367, wR2 = 0.0895
Largest diff. peak and hole / (e Å–3) 1.043 and −0.275 0.290 and −0.220
Oliveira et al. 1851Vol. 25, No. 10, 2014
309 (14267), 233 (31028); Molar conductivity (Λm in H2O):
106.80 μS cm–1; 1H NMR (DMSO, 399.8 MHz): d 2.80
(s, 6H, 2CH3C=N), 7.45 (t, 2H, J 8.0 Hz, Py-H), 7.86-7.90
(m, 6H, 2Py-H and 2NH2), 8.02 (d, 2H, J 8.0 Hz, Py-H),
8.07 (t, 2H, J 8.0 Hz, Py-H).
[Co(atc-Me)2]Cl (2)
Yield 0.085 g (67%). Analysis: Calc. for C18H22ClCoN8S2:
C, 42.48; H, 4.36; N, 22.02%. Found: C, 42.68; H, 4.55;
N, 21.61%. IR (KBr) ν/cm−1 3190, 1598, 1560, 1076,
773; UV-Vis in 1.96 × 10–5 mol L–1 MeOH solution
[λ / nm (ε / L mol–1 cm–1)]: 415 (10178), 367 (18214), 314
(24897), 260 (40102); Molar conductivity (Λm in H2O):
160.00 μS cm–1; 1H NMR (DMSO, 399.8 MHz): d 2.85
(s, 9H, 2CH3C=N and 1CH3NH), 2.97 (s, 3H, CH3NH),
7.45-8.09 (m, 8H, Py-H), 8.16 (s, 1H, NHCH3), 8.80 (s,
1H, NHCH3).
[Co(atc-Et)2]Cl·H2O(3·H2O)
Yield 0.086 g (61%). Analysis: Calc. for
C20H28ClCoN8OS2: C, 43.28; H, 5.09; N, 20.19%. Found:
C, 43.83; H, 5.17; N, 20.22%. IR (KBr) ν/cm−1 3197, 1598,
1556, 1078, 773; UV-Vis in 2.98 × 10–5 mol L–1 MeOH
solution [λ / nm (ε / L mol–1 cm–1)]: 413 (10302), 369
(18993), 315 (25241); Molar conductivity (Λm in H2O):
100 μS cm–1; 1H NMR (DMSO, 399.8 MHz): d 1.00-1.20
(m, 6H, -CH2CH3), 2.84 (s, 6H, CH3C=N), 3.20-3.43 (m,
4H, -CH2CH3), 7.45-8.08 (m, 8H, Py-H), 8.16 (s, 1H,
NHCH2CH3), 8.86 (s, 1H, NHCH2CH3).
[Co(atc-Ch)2]Cl (4)
Yield 0.086 g (53%). Analysis: Calc. for C28H38ClCoN8S2:
C, 52.13; H, 5.94; N, 17.37%. Found: C, 51.71; H, 6.70;
N, 16.81%. IR (KBr) ν/cm−1 3201, 1598, 1556, 1072, 775;
UV-Vis in 2.48 × 10–5 mol L–1 MeOH solution [λ / nm
(ε / L mol–1 cm–1)]: 428 (10967), 372 (21129), 316 (25564);
Molar conductivity (Λm in MeOH): 100 μS cm–1; 1H NMR
(DMSO, 399.8 MHz): d 1.06-1.95 (m, 20H, Ch-H), 2.83
(s, 6H, CH3C=N), 3.84 (s, 2H, CH-Ch), 7.44-8.05 (m, 8H,
Py-H), 8.86 (s, 2H, NH-Ch).
[Co(atc-Ph)2]Cl (5)
Yield 0.126 g (80%). Analysis: Calc. for C26H29ClCoN8S2:
C, 53.12; H, 4.14; N, 17.70%. Found: C, 52.80; H, 4.73; N,
17.25%. IR (KBr) ν/cm−1 3174, 1598, 1552 , 1074 , 752;
UV-Vis in 2.21 × 10–5 mol L–1 MeOH solution [λ / nm
(ε / L mol–1 cm–1)]: 385 (20588), 315 (18190), 253 (37239);
Molar conductivity (Λm in MeOH): 88 μS cm–1; 1H NMR
(DMSO, 399.8 MHz): d 2.99 (s, 6H, CH3C=N), 7.08 (t,
2H, J 6.0 Hz, Ph-H), 7.38 (t, 4H, J 8.0 Hz, Ph-H), 7.51 (t,
2H, J 6.0 Hz, Py-H), 7.69 (d, 4H, J 8.0 Hz, Ph-H), 8.02
(d, 2H, J 4.0 Hz, Py-H), 8.12-8.18 (m, 4H, Py-H), 10.37
(s, 2H, NH-Ph).
Results and Discussion
Synthesis of the complexes
Reactions of CoCl2∙6H2O with two equivalents
of Hatc-R in EtOH under reflux for 2 h results in
microcrystalline precipitates of the cobalt complexes 1-5 in
good yields (Scheme 1). Elemental analyses are consistent
with the formation of cationic complexes [Co(atc-R)2]+, in
accordance with the observed molar conductivity values.
All the compounds except [Co(atc-Ph)2]Cl are water
soluble. They are very soluble in methanol and dimethyl
sulfoxide and sparingly soluble in dichloromethane and
chloroform, demonstrating a high hydrophilic character.
Infrared, UV-Vis and 1H NMR spectroscopies
The IR spectra of the TSC ligands are characterized
by strong broad ν(NH) absorptions in the range
3365-3153 cm−1. One of them, around 3300 cm–1, is absent
Scheme 1. Synthesis of the CoIII complexes.
Cobalt(III) Complexes with Thiosemicarbazones as Potential anti-Mycobacterium tuberculosis Agents J. Braz. Chem. Soc.
1852
in the spectra of the CoIII cationic complexes, according to
the monodeprotonation of these ligands.
The ν(C=N) stretching band found around 1580 cm−1
for the free Hatc-R is observed in the 1552-1620 cm–1 range
for the complexes. The ν(N-N) band at higher frequencies
in the IR spectra of the complexes, between 1035 and
1078 cm–1, comparing to those observed for the ligands,
in the 989-995 cm–1 range, confirms coordination through
the azomethine nitrogen atom.13,34
The ν(C=S) bands appear in two regions
(1118-1074 cm–1 and 800-846 cm–1) for the free
thiosemicarbazones,35 while for the complexes the C=S
only one band is observed (752-775 cm–1), indicating
coordination through the sulfur atom and being consistent
with the deprotonation and consequent formation of a
C–S single bond in the thiosemicarbazone ligands.11 The
IR absorption bands assigned for the free ligands and
their cobalt complexes are consistent with the tridentate
coordination of the thiosemicarbazone derivatives in
a N,N,S-tridentate mode, through the sulfur atom, the
azomethine nitrogen and the pyridine nitrogen atoms,
forming octahedral complexes.
The electronic spectra of the ligands show a band in
the 312-402 nm range, assignable to a combination of
internal n→π* and π→π*electronic transitions related to
the pyridine ring.30,36-39 The spectra of the CoIII complexes
show the pyridine ring transitions, with the n→π* occurring
at higher energies, below 300 nm, confirming the complexes
formation.40 Additional bands in the 360-400 nm range
are assignable as combinations of dd transitions with
SCoIII and PyCoIII charge transfer transitions.40
Therefore, it was not possible to see the CoIII 1T1g 1A1g
and 1T2g 1A1g allowed transitions usually observed in
the visible region.41
The 1H chemical shift values of the free ligands were
previously reported.8,30,35 The CoIII complexes show similar
1H NMR behavior, with the hydrogen signals being found
as expected. The NH2 hydrogens of complex 1 are found
at 7.90 ppm as a broad singlet, while the spectra of the
complexes 4 and 5 showed the NH signals at 8.86 and
10.37 ppm, respectively. For the complexes 2 and 3, however,
two different signals relative to NH hydrogen atoms were
observed, at 8.81 and 8.16 ppm for 2 and at 8.86 and
8.17 ppm for 3. This is in accord with the non-equivalence
also observed for the methyl and ethyl groups, suggesting
that rotation around the C−NHR (R = CH3, −CH2CH3)
bond is not totally restricted.42 In fact, the methyl groups
attached to the NH fragments are also observed at different
chemical shifts for 2, one at 2.97 ppm and another one
around 2.85 ppm, overlapped with the signal for the methyl
group of the CH3−C=N moiety. The integration for this
broad peak (9 H) is consistent with this overlapping. In the
case of the complex 3, however, this observation is difficult
to confirm with certainty due to the overlapping of CH2
peak by residual water, while the methyl group appears as
a broad multiplet. For the complexes 1-5, the methyl group
(CH3−C=N) is observed in the 2.80-2.99 ppm region. The
aromatic protons are observed between 7.08 and 8.18 ppm
for all the complexes.
Crystal structures
ORTEP drawings of complexes 1∙H2O and 5∙MeOH with
numbering scheme are represented in Figure 1. Crystal data
and structure refinement for both compounds are depicted in
Table 1. The cobalt complexes are monocationic, presenting
a chloride as counter ion. The thiosemicarbazone ligands
are coordinated to the CoIII center in N,N,S-tridentate mode
Figure 1. An ORTEP view of [Co(atc)2]Cl∙H2O (1∙H2O) (left) and [Co(atc-Ph)2]Cl∙MeOH (5∙MeOH) (right). Chloride and solvate molecules were omitted
for clarity.
Oliveira et al. 1853Vol. 25, No. 10, 2014
and monoanionic form through the pyridine nitrogen atoms
N(1A) and N(1B), azomethine atoms N(2A) and N(2B) and
sulfur atoms S(1A) and S(1B).
The CoIII complexes are clearly characterized by the
smaller bond lengths compared to MnII [Mn(atc-Et)2]
compound previously reported.28 This fact is assigned to
the change of the oxidation state +II to +III, resulting in a
larger attraction of the electrons from donor atoms of the
ligand. This fact is easily observed comparing the distances
of Co(1)–N(1) and Mn–N(1) 1.952(3) e 2.2806(15) Å,
respectively. The distance Co(1)–S(1A) = 2.2035(6) Å is
also shorter compared to Mn–S(1A) = 2.5216(5) Å distance.
Furthermore, the bond distances C(8A)–S(1A), 1.745(3) and
1.745(2) Å for complexes 1∙H2O and MeOH, respectively,
are consistent with a single bond character. On the other
hand, the bond distance N(3A)–C(8A), 1.321(4) Å in 1∙H2O
and 1.314(3) Å in 5∙MeOH, shows a double bond character,
in accordance with the deprotonation of the TSCs ligands.
The coordination geometry around the metal
center is a distorted octahedron with the tridentate
thiosemicarbazone ligands perpendicular to each other
with N(1A)Co(1)N(2B) being close to 90º in both
complexes. A quite smaller distortion of the octahedral
angles is observed for the CoIII complexes when compared
with similar MnII compounds.28 This fact can be observed
through the bond angle N(2B)–M(1)–N(2A) that is around
159º in [Mn(atc-Et)2] and 178º in the CoIII complexes
studied here. The bond lengths are similar to those found for
other similar CoIII complexes.43 Selected data of interatomic
distances and main angles can be found in Table 2.
The crystal structure of 5∙MeOH is stabilized by
intermolecular hydrogen bonds, as shown in Figure 2. The
nitrogen atom N(4A) is H-bonded through H(4a) to the
oxygen atom O(1c) from a methanol molecule, while the
nitrogen atom N(4b) is H bonded with a chloride ion Cl(1),
which also interacts with a solvate molecule. The interactions
build a zigzag alignment of the species parallel to the c axis.
The crystal of 1∙H2O is built up by intermolecular
hydrogen bonds in diverse directions (see Supplementary
Information), which involve the NH2 groups, water solvate
molecules and chloride ions.
Electrochemical studies
All complexes presented a similar CV behavior,
showed exemplarily for 3 in Figure 3. One irreversible
process and two well-defined quasi-reversible (ipa/ipc 1)
waves are detected. The irreversible peak around 1.2 V is
assigned as an oxidation process involving the TSC ligand,
as previously reported for a similar compound.44 The two
cathode processes correspond to the CoIII/CoII and CoII/CoI
couples, while the two anodic processes correspond to the
CoI/CoII and CoII/CoIII couples. The complexes presented
here have an electrochemical behavior similar to that
observed for other CoIII complexes already reported.44
Through the results depicted in Table 3, it is possible
to observe the inductive effects of the R group bonded
to the N(4) atom of the thiosemicarbazone ligand on
the redox potential values. The electron donating group
(R = Ch) tends to provide the more negative potential
(E1/2 = −1.10 V) while the electron withdrawing group
Table 2. Selected bond lengths (Å) and angles (º) refined from X-ray
for 1∙H2O and 5∙MeOH
1∙H2O5∙MeOH
Bond Lengths
Co(1)–N(1A) 1.952(3) 1.9579(18)
Co(1)–N(2A) 1.893(2) 1.8861(18)
Co(1)–N(1B) 1.959(3) 1.9659(18)
Co(1)–N(2B) 1.893(2) 1.8826(18)
Co(1)–S(1A) 2.2185(10) 2.2035(6)
Co(1)–S(1B) 2.2151(10) 2.2125(6)
S(1A)–C(8A) 1.745(3) 1.745(2)
S(1B)–C(8B) 1.737(3) 1.741(2)
N(3A)–C(8A) 1.321(4) 1.314(3)
Bond Angles
N(2A)–Co(1)–N(1A) 82.82(11) 82.73(7)
N(2B)–Co(1)–N(1B) 82.44(11) 82.55(7)
N(2B)–Co(1)–N(1A) 95.85(11) 99.09(8)
N(2B)–Co(1)–N(2A) 178.54(12) 177.90(7)
N(2A)–Co(1)–N(1B) 98.15(11) 98.54(7)
N(1A)–Co(1)–N(1B) 90.23(11) 90.41(7)
N(2B)–Co(1)–S(1A) 95.43(9) 91.96(6)
C(8A)–S(1A)–Co(1) 94.58(12) 94.47(7)
N(1A)–Co(1)–S(1A) 168.72(8) 168.94(6)
Figure 2. Crystalline and molecular structure of [Co(atc-Ph)2]
Cl∙MeOH [N(4a)–H(4a)∙∙∙O(1c) = 158.3º, N(4b)–H(4)∙∙∙Cl(1) = 177.9º,
O(1)H(1)∙∙∙Cl(1) = 159.2º]. Symmetry operation used to generate O(1c),
H(1c) and Cl(1c): x, –y – 1/2, z – 1/2.
Cobalt(III) Complexes with Thiosemicarbazones as Potential anti-Mycobacterium tuberculosis Agents J. Braz. Chem. Soc.
1854
(R = Ph) shifted the process to a less negative potential
(E1/2 = −0.70 V), according to the order: −cyclohexyl <
−ethyl < −hydrogen < −methyl < −phenyl relative to
CoIII/CoII couple. In this context, the process relative to the
CoII/CoI couple presents the same trend, demonstrating
the same behavior to the first couple. In relation to the
second redox pair CoII/CoI the lower half-wave potential
is equal to −1.57 V (complex 4) and the higher potential
is E1/2 = −1.46 V (complex 5). Finally it is evidenced that
phenyl stabilizes better the oxidation state +II, while the
cyclohexyl group, with electron donating effect, reaches the
oxidation state +III easier than the other groups.
We previously reported a similar trend with MnII
complexes relative to oxidation of MnII/MnIII and
MnIII/MnIV 28 with four of those ligands. However the
influence of the groups bonded to N(4) atom observed
for MnII complexes is different for each redox pair. By
comparing the values found for manganese complexes,
it is possible to conclude that cobalt compounds oxidize
much more easily than the manganese ones, since cobalt
complexes are stabilized in oxidation state +III.
Biological activity
The biological activity of the compounds was verified
by determining the values of MIC against strains of
Mycobacterium tuberculosis H37Rv ATCC 27294. Synthetic
compounds with MIC 12.5 μg mL–1 are considered of
interest to be further evaluated in cytotoxicity tests, which
were primarily evaluated using normal epithelial cells
(VERO). Complex 5, with SI 10 (SI = IC50/MIC) for
VERO cells, was further analyzed against macrophages
cells J774A.1 (immunologic system cells).
The biological results (anti-MTB activity and cytotoxicity
against VERO cells) are shown in Table 4. Two cobalt
complexes present MIC 12.50 μg mL–1, 4 and 5, with
values of 2.41 μmol L–1 and 9.87 μmol L–1, respectively.
Complex 4 presented a similar activity as the free Hatc-Ch
ligand (MIC = 2.82 μmol L–1)27 while complex 5 was more
active than the free Hatc-Ph ligand (MIC = 57.75 μmol L–1),27
in this case improving the activity by complexation. In
the other cases the complexation to the CoIII didn’t lead
to improvement on the activities in relation to the free
ligands. The cobalt salt CoCl26H2O was not effectively
active (MIC > 105 μmol L–1) showing that the activity of the
complexes cannot be merely associated to the presence of
the metal ion. Complex 5 presented quite low cytotoxicity
against VERO cells and therefore was also investigated on
macrophages cell line J774A.1 (IC50 = 988.79 μmol L–1)
resulting in high selectivity (SI = 100).
NiII and MnII structurally related compounds studied
before14,28 showed to be more active in vitro than the
CoIII analogs studied here. This fact can be explained
by the increased polarity of the ionic CoIII compounds
compared with the neutral NiII and MnII complexes, which
can influence the permeability through the lipid layer of
bacterial membrane resulting in a lower cellular inflow
of the active species.45 Otherwise, the cationic cobalt
complexes are very selective and also more water soluble
than the neutral nickel or manganese species, which could
enhance their absorption in vivo, compensating the eventual
lower cellular permeation.
The high SI found for complex 5 shows its potential
for clinical use, with a wide difference between the
concentrations regarding the activity on the pathogen
and the cytotoxicity on normal epithelial VERO cells,
respectively. Furthermore, at the concentration the complex
is active on the pathogen it remains innocuous front the
Figure 3. Cyclic voltammogram of [Co(atc-Et)2]+ (scan rate 100 mV s–l)
full amplitude and narrow amplitudes.
Table 3. Cyclic voltammetry for the redox couples CoIII/CoII and CoII/CoI for all four complexes, measured in acetonitrile with 0.1 M PTBA as the electrolyte
Complex ECoIII/CoII ECoII/CoIII E1/2 / V ECoII/CoIECoI/CoII E1/2 /V
1−0.86 −0.76 −0.80 −1.54 −1.50 −1.52
2−0.80 −0.73 −0.77 −1.54 −1.47 −1.50
3−0.86 −0.75 −0.80 −1.60 −1.48 −1.54
4−1.15 −1.02 −1.10 −1.60 −1.54 −1.57
5−0.73 −0.65 −0.70 −1.52 −1.41 −1.46
Oliveira et al. 1855Vol. 25, No. 10, 2014
Table 4. Anti-MTB activity (MIC), cytotoxicity (IC50), and selectivity index (SI) of the complexes
Compound MIC IC50 (VERO cells) SIa
/ μg mL–1 / μmol L–1 / μg mL–1 / μmol L–1
[Co(atc)2]Cl∙H2O (1∙H2O) > 25 > 50 434.30 872.05
[Co(atc-Me)2]Cl (2) > 25 > 49 483.50 950.69
[Co(atc-Et)2]Cl (3) > 25 > 46 95.90 178.89
[Co(atc-Ch)2]Cl (4) 1.56 2.41 11.80 18.31 7.5
[Co(atc-Ph)2]Cl (5) 6.25 9.87 105.30 166.59 17
CoCl26H2O > 25 > 105 368.40 3409.2
Isoniazid 0.03 0.21
aThe selectivity index (SI) was calculated by the ratio IC50VERO/MIC. MIC values for the ligands: Hatc = 31.3 μg mL–1 (161.1 μmol L–1); Hatc-Me = 7.8 μg mL–1
(37.4 μmol L–1); Hatc-Et = 3.13 μg mL–1 (14.08 μmol L–1); Hatc-Ch = 0.78 μg mL–1 (2.82 μmol L–1); Hatc-Ph = 15.6 μg mL–1 (57.75 μmol L–1).27
macrophage cells (J774A.1), which represent the first
immune response to the infection.
Conclusions
A series of CoIII compounds with thiosemicarbazones
ligands, changing the N(4) substituent group by H,
methyl, ethyl, cyclohexyl and phenyl could be synthesized
in satisfactory yields and fully characterized. The
coordination of Hatc-Ph to the CoIII was able to enhance
the anti-M. tuberculosis activity when compared with the
free ligand. Moreover the results confirm that structural
changes in peripheral group of the ligands can affect
significantly the activity against M. tuberculosis as well
as the cytotoxicity. Complex 5 showed potential biological
results, not being the most active on the pathogen, but acting
more selectively and thus showing higher potentiality as
anti-M. tuberculosis agent.
Supplementary Information
Supplementary data are available free of charge at
http://jbcs.sbq.org.br as PDF file.
Acknowledgments
The authors thank FAPESP (Grants 2009/54011-8,
2011/16380-1 and 2013/14957-5), CNPq and CAPES for
supporting this work. This work is also a collaboration
research project of a member of the Rede Mineira de
Química (RQ-MG) supported by FAPEMIG (Project:
REDE-113/10).
References
1. http://www.who.int/tb, accessed in June 2014.
2. Li, R.; Sirawaraporn, R.; Chitnumsub, P.; Sirawaraporn, W.;
Wooden, J.; Athappilly, F.; Turley, S.; Hol, W. G. J.; J. Mol.
Biol. 2000, 295, 307.
3. Ananthan, S.; Faaleolea, E. R.; Goldman, R. C.; Hobrath,
J. V.; Kwong, C. D.; Laughon, B. E.; Maddry, J. A.; Mehta, A.;
Rasmussen, L.; Reynolds, R. C.; Secrist III, J. A.; Shindo, N.;
Showe, D. N.; Sosa, M. I.; Suling, W. J.; White, E. L.;
Tuberculosis 2009, 89, 334.
4. Lin, P. L.; Dartois, V.; Johnston, P. J.; Janssen, C.; Via, L.;
Goodwin, M. B.; Klein, E.; Barry, C. E.; Flynn, J. L.; PNAS
2012, 109, 14188.
5. Lun, S.; Guo, H.; Onajole, O. K.; Pieroni, M.; Gunosewoyo, H.;
Chen, G.; Tipparaju, S. K.; Ammerman, N. C.; Kozikowski, A. P.;
Bishai, W. R.; Nat. Commun. 2013, 4, 2907.
6. Hong, X.; Hopfinger, A. J.; Biomacromolecules 2004, 5,
1066.
7. Lång, H.; Quaglio, G.; Olesen, O. F.; Tuberculosis 2010,
90, 1.
8. Maia, P. I. S.; Pavan, F. R.; Leite, C. Q. F.; Lemos, S. S.; de Sousa,
G. F.; Batista, A. A.; Nascimento, O. R.; Ellena, J.; Castellano,
E. E.; Niquet, E.; Deflon, V. M.; Polyhedron 2009, 28, 398.
9. Tarallo, M. B.; Urquiola, C.; Monge, A.; Costa, B. P.; Ribeiro,
R. R.; Costa-Filho, A. J.; Mercader, R. C.; Pavan, F. R.; Leite,
C. Q. F.; Torre, M. H.; Gambino, D.; J. Inorg. Biochem. 2010,
104, 1164.
10. Alessio, E.; Bioinorganic Medicinal Chemistry, 1st ed.;
WILEY-VCH: Trieste, Italy, 2011.
11. Maia, P. I. S.; Graminha, A.; Pavan, F. R.; Leite, C. Q. F.;
Batista, A. A.; Back, D. F.; Lang, E. S.; Ellena, J.; Lemos, S. S.;
Salistre-de-Araujo, H. S.; Deflon, V. M.; J. Braz. Chem. Soc.
2010, 21, 1177.
12. Chen, C.; Zhu, X.; Li, M.; Guo, H.; Niu, J.; Russ. J. Coord.
Chem. 2011, 37, 435.
13. Batista, D. da G. J.; Silva, P. B.; Lachter, D. R.; Silva, R. S.;
Aucelio, R. Q.; Louro, S. R. W.; Beraldo, H.; Soeiro, M. N. C.;
Teixeira, L. R.; Polyhedron 2010, 29, 2232.
Cobalt(III) Complexes with Thiosemicarbazones as Potential anti-Mycobacterium tuberculosis Agents J. Braz. Chem. Soc.
1856
14. Maia, P. I. S.; Pavan, F. R.; Leite, C. Q. F.; Abram, U.; Lang,
E. S.; Batista, A. A.; Deflon, V. M. In Metal Ions in Biology and
Medicine and Aqueous Chemistry and Biochemistry of Silicon.
John Libbey Eurotext: Paris, 2011, p. 164–171.
15. Bertini, I.; Gray, H. B.; Lippard, S. J.; Valentine, J. S.;
Bioinorganic Chemistry, 1st ed.; University Science Books Mill
Valley: California, USA, 1994.
16. García-Tojal, J.; García-Orad, A.; Díaz, A. A.; Serra, J. L.;
Urtiaga, M. K.; Arriortua, M. I.; Rojo, T.; J. Inorg. Biochem.
2001, 84, 271.
17. Fan, X.; Dong, J.; Min, R.; Chen, Y.; Yi, X.; Zhou, J.; Zhang, S.;
J. Coord. Chem. 2013, 66, 4268.
18. Ramachandran, E.; Thomas, S. P.; Poornima, P.; Kalaivani, P.;
Prabhakaran, R.; Padma, V. V.; Natarajan, K.; Eur. J. Med.
Chem. 2012, 50, 405.
19. Mondelli, M.; Pavan, F. R.; Souza, P. C.; Leite, C. Q. L.;
Ellena, J.; Nascimento, O. R.; Facchin, G.; Torre, M. H.; J. Mol.
Struct. 2013, 1036, 180.
20. Hoffman, A. E.; DeStefano, M.; Shoen, C.; Gopinath, K.;
Warner, D. F.; Cynamon, M.; Doyle, R. P.; Eur. J. Med. Chem.
2013, 70, 589.
21. Maia, P. I. S.; Nguyen, H. H.; Hagenbach, A.; Bergemann, S.;
Gust, R.; Deflon, V. M.; Abram, U.; Dalton Trans. 2013, 42,
5111.
22. García-Gallego, S.; Serramía, M. J.; Arnaiz, E.; Díaz, L.;
Muñoz-Fernández, M. A.; Gómez-Sal, P.; Ottaviani, M. F.;
Gómez, R.; Javier de la Mata, F.; Eur. J. Inorg. Chem. 2011,
1657.
23. Santos, D.; Parajón-Costa, B.; Rossi, M.; Caruso, F.; Benítez, D.;
Varela, J.; Cerecetto, H.; González, M.; Gómez, N.; Caputto,
M. E.; Moglioni, A. G.; Moltrasio, G. Y.; Finkielsztein, L. M.;
Gambino, D.; J. Inorg. Biochem. 2012, 117, 270.
24. Klayman, D. L.; Bartosevich, J. F.; Griffin, T. S.; Mason, C. J.;
Scovill, J. P.; J. Med. Chem. 1979, 22, 855.
25. Stefani, C.; Punnia-Moorthy, G.; Lovejoy, D. B.; Jansson, P. J.;
Kalinowski, D. S.; Sharpe, P. C.; Bernhardt, P. V.; Richardson,
D. R.; J. Med. Chem. 2011, 54, 6936.
26. Soares, M. A.; Lessa, J. A.; Mendes, I. C.; Silva, J. G.; Santos,
R. G.; Salum; L. B.; Daghestani, H.; Andricopulo, A. D.; Day,
B. W.; Vogt, A.; Pesquero, J. L.; Rocha, W. R.; Beraldo, H.;
Bioorg. Med. Chem. 2012, 20, 3396.
27. Pavan, F. R.; Maia, P. I. S.; Leite, S. R. A.; Deflon, V. M.;
Batista, A. A.; Sato, D. N.; Franzblau, S. G.; Leite, C. Q. F.;
Eur. J. Inorg. Chem. 2010, 45, 1898.
28. Oliveira, C. G.; Maia, P. I. S.; Souza, P. C.; Pavan, F. R.; Leite,
C. Q. F.; Viana, R. B.; Batista, A. A.; Nascimento, O. R.; Deflon,
V. M.; J. Inorg. Biochem. 2014, 132, 21.
29. Richardson, D. R.; Kalinowski, D. S.; Richardson, V.; Sharpe,
P. C.; Lovejoy, D. B.; Islam, M.; Bernhardt, P. V.; J. Med. Chem.
2009, 52, 1459.
30. West, D. X.; Billeh, I. S.; Jasinski, J. P.; Jasinski, J. M.; Butcher,
R. J.; Transit. Metal Chem. 1998, 23, 209.
31. Sheldrick, G. M.; SHELXS97; Program for the Solution of
Crystal Structures; University of Göttingen, Germany, 1997.
32. Sheldrick, G. M.; SHELXL97; Program for the Renement of
Crystal Structures; University of Göttingen, Germany, 1997.
33. Palomino, J. C.; Martin, A.; Camacho, M.; Guerra, H.;
Swings, J.; Portaels, F.; Antimicrob. Agents Chemother. 2002,
46, 2720.
34. Maia, P. I. S.; Nguyen, H. H.; Ponader, D.; Hagenbach, A.;
Bergemann, S.; Gust, R.; Deflon, V. M.; Abram, U.; Inorg.
Chem. 2012, 51, 1604.
35. Wiles, D. M.; Gingras, B. A.; Suprunchuk, T.; Can. J. Chem.
1967, 45, 469.
36. Chan, J.; Thompson, A. L.; Jones, M. W.; Peach, J. M.; Inorg.
Chim. Acta 2010, 363, 1140.
37. Enyedy, E. A.; Primik, M. F.; Kowol, C. R.; Arion, V. B.;
Kiss, T.; Keppler, B.; Dalton Trans. 2011, 40, 5895.
38. West, D. X.; Beraldo, H.; Nassar, A. A.; El-Saied, F. A.; Ayad,
M. I.; Transition Met. Chem. 1999, 24, 595.
39. Pavia, D. L.; Lampman, G. M.; Kriz, G. S.; Introduction to
Spectroscopy. A Guide for Students of Organic Chemistry; 3rd
ed.; Bellingham, USA, 2001.
40. West, D. X.; Lockwood, M. A.; Transition Met. Chem. 1997,
22, 447.
41. Housecroft, C. E.; Sharpe, A. G. Inorganic Chemistry; 2nd ed.;
Prentice Hall: Harlow, England, 2005.
42. Nguyen, H. H.; Jegathesh, J. J.; Maia, P. I. S.; Deflon, V. M.;
Gust, R.; Bergemann, S.; Abram, U.; Inorg. Chem. 2009, 48,
9356.
43. Zhu, X. F.; Fan, Y. H.; Wang, Q.; Chen, C. L.; Li, M. X.; Zhao,
J. W.; Zhou, J.; Russ. J. Coord. Chem. 2012, 38, 478.
44. Rodić, M. V.; Leovac, V. M.; Jovanović, L. S.; Vojinović-Ješić,
L. S.; Divjaković, V.; Češljević, V. I.; Polyhedron 2012, 46, 124.
45. Maurya, M. R.; Kumar, A.; Abid, M.; Azam, A.; Inorg. Chim.
Acta 2006, 359, 2439.
Submitted: March 31, 2014
Published online: June 25, 2014
FAPESP has sponsored the publication of this article.
Supplementary Information SI
J. Braz. Chem. Soc., Vol. 25, No. 10, S1-S6, 2014.
Printed in Brazil - ©2014 Sociedade Brasileira de Química
0103 - 5053 $6.00+0.00
*e-mail: deflon@iqsc.usp.br
Cobalt(III) Complexes with Thiosemicarbazones as Potential
anti‑Mycobacterium tuberculosis Agents
Carolina G. Oliveira,a Pedro Ivo da S. Maia,b Marcelo Miyata,c Fernando R. Pavan,c
Clarice Q. F. Leite,c Eduardo Tonon de Almeidad and Victor M. Deflon*,a
aInstituto de Química de São Carlos, Universidade de São Paulo, 13566-590 São Carlos-SP, Brazil
bDepartamento de Química, Universidade Federal do Triângulo Mineiro, 38025-440 Uberaba-MG, Brazil
cFaculdade de Ciências Farmacêuticas, Universidade Estadual Paulista, 14801-902 Araraquara-SP, Brazil
dInstituto de Química, Universidade Federal de Alfenas, 37130-000 Alfenas-MG, Brazil
Crystallographic data (excluding structure factors) for
the structures in this work were deposited in the Cambridge
Crystallographic Data Centre as supplementary publication
numbers CCDC 988971 (1·H2O) and 988972 (5·MeOH).
Copies of the data can be obtained, free of charge, via
www.ccdc.cam.ac.uk/conts/retrieving.html or from the
Cambridge Crystallographic Data Centre, CCDC, 12 Union
Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.
E-mail: deposit@ccdc.cam.ac.uk.
IR spectra of compounds
Figure S1. IR spectrum (KBr) of complex 1.
Cobalt(III) Complexes with Thiosemicarbazones as Potential anti-Mycobacterium tuberculosis Agents J. Braz. Chem. Soc.
S2
Figure S2. IR spectrum (KBr) of complex 2.
Figure S3. IR spectrum (KBr) of complex 3.
Figure S4. IR spectrum (KBr) of complex 4.
Oliveira et al. S3Vol. 25, No. 10, 2014
Figure S5. IR spectrum (KBr) of complex 5.
Figure S6. 1H NMR spectrum (DMSO, 399.8 MHz) of complex 1.
1H NMR spectra of compounds
Cobalt(III) Complexes with Thiosemicarbazones as Potential anti-Mycobacterium tuberculosis Agents J. Braz. Chem. Soc.
S4
Figure S7. 1H NMR spectrum (DMSO, 399.8 MHz) of complex 2.
Figure S8. 1H NMR spectrum (DMSO, 399.8 MHz) of complex 3.
Oliveira et al. S5Vol. 25, No. 10, 2014
Figure S9. 1H NMR spectrum (DMSO, 399.8 MHz) of complex 4.
Figure S10. 1H NMR spectrum (DMSO, 399.8 MHz) of complex 5.
Cobalt(III) Complexes with Thiosemicarbazones as Potential anti-Mycobacterium tuberculosis Agents J. Braz. Chem. Soc.
S6
Figure S11. Crystalline and molecular structure of [Co(atc)2]Cl·H2O [N(24)···O(1L) = 2.910(5) Å, N(24)-H(24A)···O(1L) = 146.5º], [N(24)···Cl(1J)
= 3.258(3) Å, N(24)-H(24B)···Cl(1J) = 157.6º], [N(14)···Cl(1G) = 3.223(3)Å, N(14)-H(14A)···Cl(1G) = 147.5º], [N(14)···Cl(1H) = 3.222(3)Å, N(14)-
H(14B)···Cl(1H) = 152.3º], [O(1)···Cl(1F) = 3.081(4) Å, O(1)-H(1w)···Cl(1F) = 174(4)º]. Symmetry operations used: F 1−x,−y, 1−z; L 1/2+x,1/2−y,1/2+z
; J 1/2+x,1/2−y, −1/2+z; G 1/2−x,1/2+y, 1/2−z; H −1/2+x,1/2−y, −1/2+z.
... The compounds studied here were previously synthesized and characterized as described by Fernandes and coworkers [34]. L 1 was prepared by refluxing 4-phenyl-3-thiosemicarbazide and 2-acetylpyridine (1:1) in ethanolic solutions [35]. ...
... Briefly, the reaction of CoCl 2 •6H 2 O and L 1 (1:2) in ethanol (15 mL) provided [Co III (L 1 ) 2 ]Cl (Fig. 1). The products were characterized and purities evaluated [33,34]. The characterization of the compounds was previously certified by techniques such as 1H and 13C NMR, high-resolution mass spectrometry, LC-MS/MS and fragmentation study, and previously reported [34]. ...
... The products were characterized and purities evaluated [33,34]. The characterization of the compounds was previously certified by techniques such as 1H and 13C NMR, high-resolution mass spectrometry, LC-MS/MS and fragmentation study, and previously reported [34]. The compounds were dissolved in dimethyl sulfoxide (DMSO) and stored at − 20 °C for the biological assays. ...
Article
Full-text available
Chikungunya virus (CHIKV) is the causative agent of chikungunya fever, a disease that can result in disability. Until now, there is no antiviral treatment against CHIKV, demonstrating that there is a need for development of new drugs. Studies have shown that thiosemicarbazones and their metal complexes possess biological activities, and their synthesis is simple, clean, versatile, and results in high yields. Here, we evaluated the mechanism of action (MOA) of a cobalt(III) thiosemicarbazone complex named [CoIII(L1)2]Cl based on its in vitro potent antiviral activity against CHIKV previously evaluated (80% of inhibition on replication). Furthermore, the complex has no toxicity in healthy cells, as confirmed by infecting BHK-21 cells with CHIKV-nanoluciferase in the presence of the compound, showing that [CoIII(L1)2]Cl inhibited CHIKV infection with the selective index of 3.26. [CoIII(L1)2]Cl presented a post-entry effect on viral replication, emphasized by the strong interaction of [CoIII(L1)2]Cl with CHIKV non-structural protein 4 (nsP4) in the microscale thermophoresis assay, suggesting a potential mode of action of this compound against CHIKV. Moreover, in silico analyses by molecular docking demonstrated potential interaction of [CoIII(L1)2]Cl with nsP4 through hydrogen bonds, hydrophobic and electrostatic interactions. The evaluation of ADME-Tox properties showed that [CoIII(L1)2]Cl presents appropriate lipophilicity, good human intestinal absorption, and has no toxicological effect as irritant, mutagenic, reproductive, and tumorigenic side effects.
... Recently, we have been described Pt(II), Pd(II), Au(III), Co(III), Ni (II), Cu(II) and Mn(II) complexes containing thiosemicarbazones with anti-bacterial, anti-viral, antiparasitic and anti-cancer activity [11][12][13][14][15]. As a continuation of our research in the search for new active compounds, and based on the fact that zinc is the second most abundant dblock metal in the body, being found in bones, muscles and different organs of the human body [16], we describe here, the synthesis and characterization of new three zinc(II) complexes, including crystal structure, Hirshfeld surface and spectroscopy analysis. ...
... The electronic spectra of the free ligands present a single band in the 250-350 nm range, probably associated to a combination of internal n → π* and π → π*electronic transitions ( Fig. S17-S21) [14]. In the electronic spectra of the synthesized complexes, additional bands in the 360-400 nm range, (Fig. 1a) are observed. ...
... In the electronic spectra of the synthesized complexes, additional bands in the 360-400 nm range, (Fig. 1a) are observed. The low energy bands can be attributed as as combinations of charge transfer transitions, as noted for analogous complexes in the literature [14]. In this case, d-d transitions are not expected in Zn(II) complexes as the metal ion has a d 10 configuration. ...
Article
In the present work, the synthesis, characterization, antifungal activity, molecular docking study and in silico approach of five thiosemicarbazone derivatives and their corresponding zinc(II) complexes are described. The compounds were characterized by elemental analysis, IR, UV–Vis and NMR spectroscopic measurements, molar conductivity measurements, emission spectra, high-resolution mass spectrometry and X ray study. The antifungal activity of the free ligands and synthesized compounds was preliminarily evaluated against Candida albicans (ATCC 90028), Candida tropicalis (ATCC 13803) and Candida glabrata (ATCC 2001), by the minimum inhibitory concentration (MIC) assay. Two complexes, 4 (MIC = 3.18 to 6.37 μM) and 5 (MIC = 25.95 μM for all) showed promising results, being highly active against all strains evaluated. The X-ray analyses shown that the complex 2 crystallizes in the centrosymmetric space group P21/c of the monoclinic system and the coordination sphere around zinc(II) atom is better described as slightly distorted octahedral. The Hirshfeld surface (HS) analysis showed that non-classical H···H and C···H/H···C contacts contribute with 65.9% while the S···H and N···H (21%) and Cl···H and O···H interactions (12%) complete the HS area. The molecular docking results, performed against CYP51 enzyme (sterol 14α-demethylase) of C. albicans and C. glabrata shows that the complexes 4 (ΔG = −10.75 and − 12.90 kcal/ mol) and 5 (ΔG = −11.12 and − 14.53 kcal/ mol) showed the highest binding free energies of all compounds. The ADME-Tox (absorption, distribution, metabolism, excretion and toxicity) in silico parameters evaluated showed promising results for all compounds.
... In some cases, the biological approaches of the TSCs can be potentialized by forming chelates with metal ions [20]. Previous studies showed that analogue octahedral Co(III), Ni(II), and Mn(II) complexes containing thiosemicarbazones derived from 2-acetylpyridine possess remarkable antibacterial activity [22][23][24][25]. Ribonucleotide reductases, enzymes responsible for catalysing a crucial step for DNA synthesis, are considered one of the main targets of TSCs [26]. ...
... The absorption spectra in the IR region of the complexes are marked by significant displacements in the wave numbers of stretches ν(C=S), ν(C=C + C-N) and ν(N-H) when compared to the free ligand. The spectral data showed that the thiosemicarbazone behaves as an NNStridentate ligand through the sulfur atoms, azomethine nitrogen and pyridine nitrogen atom [16,23,53]. The band ν(C=S) moves from 833 cm − 1 in the spectrum of the free ligand to a frequency of 782 cm − 1 in IR of the complex [CuCl(atc-Me)] (complex 1), indicating coordination via the sulfur atom. ...
... Other papers have associated the biological activity of copper complexes on the bond formed between the metal ion with nucleic acids, causing them DNA damage [8,68]. Additional changes on the peripheral part of the thiosemicarbazone could be made to acquire improved antibacterial agents [16,23]. As observed, a phenyl group substituent normally provides lower MIC values of antibacterial activity than methyl or ethyl groups [13,16]. ...
Article
Considering the promising previous results on the remarkable activity exhibited by cobalt(III) and manganese(II) thiosemicarbazone compounds as antibacterial agents, the present study aimed to prepare and then evaluate the antibacterial activity of two different types of Cu(II) complexes based on a 2-acetylpyridine-N(4)-methyl-thiosemicarbazone ligand (Hatc-Me), a monomer complex [CuCl(atc-Me)] and a novel dinuclear complex [{Cu(μ-atc-Me)}2μ-SO4]. The compounds were characterized by infrared spectra, ultraviolet visible and CHN elemental analysis. In addition, the crystalline structures of the complexes were determined by single-crystal X-ray diffraction. In both cases, the Schiff base ligand coordinated in a tridentate mode via the pyridine nitrogen, imine nitrogen and sulfur atoms. The two Cu(II) atoms in the dimer are five coordinate, consisting of three NNS-donor atoms from the thiosemicarbazone ligand connected by a sulfate bridge. The Hirshfeld surface and energy framework of the complexes were additionally analyzed to verify the intermolecular interactions. The biological activity of the Cu(II) salts, the free ligand and its Cu(II) complexes was evaluated against six strains of mycobacteria including Mycobacterium tuberculosis. The complexes showed promising results as antibacterial agents for M. avium and M. tuberculosis, which ranged from 6.12 to 12.73 μM. Furthermore, molecular docking analysis was performed and the binding energy of the docked compound [{Cu(μ-atc-Me)}2μ-SO4] with M. tuberculosis and M. avium strains were extremely favorable (−11.11 and − 14.03 kcal/mol, respectively). The in silico results show that the complexes are potential candidates for the development of new antimycobacterial drugs.
... Additionally, the activity of Co(III) with Schiff base 2-acetyl-piridine-N(4)-R-thiosemicarbazone (R = hydrogen, methyl, ethyl, cyclohexyl and phenyl) has been described in the literature. In addition to the satisfactory antibacterial activity of this type of complex on Mycobacterium tuberculosis H37Rv (ATCC 27294) bacteria with a minimal inhibitory concentration (MIC) of less than 10 µg/mL [27], the toxicity evaluated on normal cells (VERO and J774A.1) discovered the phenyl derivative complex as the highest selective agent (selective index of 17). ...
... Hatc-Me presented considerable MIC and MBC values (<12.5 µg/mL) in three tested strains, while its cobalt(III) complex (1) derivate has not shown relevant biological activity (MIC and MCB = 400 µg/mL). Surprisingly, the phenyl containing ligand Hatc-Ph did not present activity, but complex 2 showed promising activity against all the investigated bacteria, especially for S. sanguinis strain in which the MIC and MCB are equal to 0.39 µg/mL, showing bactericidal and bacteriostatic effects, and, in this case improving the activity by complexation, as previously observed [3,13,27]. Neutral nickel [37] and manganese [13] To investigate even further the antimicrobial potential of the complexes, time-kill curve assays were performed against six bacteria strains of complex 2, the most active compound. The curves were drawn using minimal bactericidal concentration (MBC) results obtained for complex 2 and are shown in Figure 3. used as positive control also presented the same bactericidal effected in less than 24 h, but the only strain that the antimicrobial eliminated faster than the compound was L. paracasei, which suggests that complex 2 has antibacterial action that is as fast as chlorhexidine. ...
... The complexes [Co(atc-Me) 2 ]Cl (1) and [Co(atc-Ph) 2 ]Cl (2) were synthesized by slight modifications of the reported procedures described by Oliveira et al. [27] An amount of 0.25 mmol of CoCl 2 ·6H 2 O was added to 0.5 mmol of Hatc-R in 10 mL ethanol. A brown solid was formed, separated by filtration and washed with n-hexane. ...
Article
Full-text available
Considering our previous findings on the remarkable activity exhibited by cobalt(III) with 2-acetylpyridine-N(4)-R-thiosemicarbazone (Hatc-R) compounds against Mycobacterium tuberculosis, the present study aimed to explored new structure features of the complexes of the type [Co(atc–R)2]Cl, where R = methyl (Me, 1) or phenyl (Ph, 2) (13C NMR, high-resolution mass spectrometry, LC–MS/MS, fragmentation study) together with its antibacterial and antiviral biological activities. The minimal inhibitory and minimal bactericidal concentrations (MIC and MBC) were determined, as well as the antiviral potential of the complexes on chikungunya virus (CHIKV) infection in vitro and cell viability. [Co(atc-Ph)2]Cl revealed promising MIC and MBC values which ranged from 0.39 to 0.78 µg/mL in two strains tested and presented high potential against CHIKV by reducing viral replication by up to 80%. The results showed that the biological activity is strongly influenced by the peripheral substituent groups at the N(4) position of the atc-R1− ligands. In addition, molecular docking analysis was performed. The relative binding energy of the docked compound with five bacteria strains was found in the range of −3.45 and −9.55 kcal/mol. Thus, this work highlights the good potential of cobalt(III) complexes and provide support for future studies on this molecule aiming at its antibacterial and antiviral therapeutic application.
... The potent chelator feature of TSCs is based on their chemical versatility, which provides variable binding modes and structural diversity that depends on the covalent sites number [10]. The ability of TSCs to form new metal complexes is well demonstrated in the literature with plenty of metal ions, including manganese(II) [11], copper(II) [12], zinc(II) [13], cobalt(III) [14], palladium(II) [15] and platinum(II) [16,17]. ...
... In the previous work, we investigated the structures and biological properties of some thiosemicarbazonate ligands and their Mn(II) [11], Co(III) [14] and Zn(II) [13] complexes. In the present article, a new synthetic route is described to obtain ionic Fe(III) thiosemicarbazonate compounds. ...
... The N,N,S coordination mode in monoanionic form of both ligands was confirmed due to significant alterations in ν(C=S), ν (C=C + C=N) and ν(N-H) stretches when compared with the free ligands. The same changes have been noted for similar complexes derived from thiosemicarbazonate moieties [11,14,15]. In the spectra of the complexes 1·2H 2 O and 2·H 2 O, the absorption related to the ν(N-H) around 3300 cm −1 was not observed, as expected by the monodeprotonation of the ligands after complexation. ...
Article
Full-text available
Reactions of FeSO4 precursor with thiosemicarbazones Hatc-R, where R is ethyl (Et) or phenyl (Ph), led to the formation of homoleptic iron(III) complexes of the type [Fe(atc-R)2]HSO4. The characterization of the compounds was performed by spectroscopy techniques, such as FTIR, UV–Vis, besides elemental analysis, conductometry, voltammetry and magnetic susceptibility measurement. The crystalline structure of [Fe(atc-Ph)2]HSO4·H2O was determined by single-crystal X-ray diffraction and revealed the oxidation of the Fe(II) centre to Fe(III) upon complexation of the monoanionic N,N,S-tridentate thiosemicarbazonate ligands. The magnetic susceptibility results showed the paramagnetic property of the iron(III) complexes in the extension of 1 unpaired electron. The electrochemical analyses showed a nearly reversible process of the iron complex, which is slightly influenced by the peripheral substituent groups at the N(4) position of the atc-R1− ligands. Hirshfeld surface analysis revealed that the supramolecular structure of [Fe(atc-Ph)2]HSO4·H2O is stabilized mainly by H···H, C···H/H···C and O···H/H···O interactions.
... Quercetin has poor solubility, so it seems to be difficult to absorb into the body [17,18]. Several studies have been performed to modify the quercetin structure to increase its water solubility and bioavailability and thus enhance its pharmacological effects [19][20][21][22][23]. Studies have shown that coordination of quercetin with metal ions can increase the antioxidant activity and ultraoxygen anion elimination than quercetin itself [24][25][26]. e strong ability of quercetin to chelate with different metal ions such as Tb(III) [27], Mg(II) [28], Cu(II) [29], Fe(II) [3], Cr(III) [30], Co(II) [31], Sn(II) [32], Vo(IV) [33], Zn(II) [34], Mn(II) [35], Pb(II) [36], and Ni(II) [37] can increase the solubility and bioavailability of quercetin and promote new pharmacological activity [38]. ...
Article
Full-text available
Quercetin (3,3′,4′,5,7-pentahydroxyflavone) is one of the dietary flavonoids, distributed in medicinal plants, vegetables, and fruits. Quercetin has the ability to bind with several metal ions to increase its biological activities. In the last two decades, quercetin has attracted considerable attention due to the biological and pharmaceutical activities such as antioxidant, antibacterial, and anticancer. In the present study, quercetin and ethanolamine were used for the synthesis Schiff base complex, which was characterized by IR, ¹H NMR, and ¹³C NMR spectroscopy. The Schiff base has been employed as a ligand for the synthesis of novel nanoscale Cu (II) complex. The product was characterized by FT-IR spectroscopy, FESEM, and XRD. Significantly, the product showed remarkable catalytic activity towards the oxidation of primary and secondary alcohols. The antibacterial activity of the final product was assessed against Staphylococcus aureus (Gram‐positive) and Escherichia coli (Gram‐negative) bacteria using an inhibition zone test. The synthesized nanoscale Cu (II) complex exhibited a strong antibacterial activity against both Gram-positive and Gram-negative bacteria. 1. Introduction Flavonoids are the largest group of phenolic compounds that have different biological and medicinal properties, such as antioxidant, [1] antibacterial, [2] antidiabetic, [3] anticancer, [4–6] antiatherosclerosis [7], and neuroprotective effects [8]. Flavonoids consist of two benzene rings joined by a 3-carbon bridge (C6-C3-C6) (Figure 1) [9]. Flavonoids can be divided into several different classes, such as flavones (e.g., flavone, luteolin, and apigenin), flavonols (e.g., quercetin, kaempferol, fisetin, and myricetin), and flavanones (e.g., flavanone, naringenin, and hesperetin). These classes are different in the oxidation and pattern of substitution of the C ring, while in each class, they differ in the pattern of substitution of the A and B rings [10].
... Since 1946, the activity of thiosemicarbazone against M. tuberculosis has been demonstrated [8][9][10]. Additionally, a previous study conducted in our laboratory, evaluated the in vitro anti-M. ...
Article
Background For more than 60 years, the lack of new anti-tuberculosis drugs and the increase of resistant Mycobacterium tuberculosis lineages exhibit a therapeutic challenge, demanding new options for the treatment of resistant tuberculosis. Objective Herein, we determined the (i) activities of (-)-camphene and derivatives and (ii) combinatory effect with pyrazinamide (PZA) against Mycobacterium tuberculosis in acidic pH and (iii) cytotoxicity in VERO cells. Methods The activity of (-)-camphene and 15 derivatives were determined in M. tuberculosis H37Rv in culture medium at pH 6.0 by Resazurin Microtiter Assay Plate (REMA). The combinatory study of three (-)-camphene derivatives with PZA was carried out in seven multidrug-resistant (MDR) clinical isolates by REMA and Checkerboard, respectively. The assay of 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium (MTT) bromide in VERO cells was used to determine the derivatives cytotoxicity. Results Four (-)-Camphene derivatives, (4), (5a) (5d) and (5h), showed reduction in MIC value at pH 6.0 compared to MIC detected at pH 6.8 in M. tuberculosis H37Rv and multidrug resistant clinical isolates. Three (-)-camphene derivatives, (4), (5d) and (5h), showed synergistic effect (FICI ≤ 0.5) combined with PZA and were more selective for M. tuberculosis than VERO cell (selective index from 7.7 to 84.2). Conclusion Three (-)-camphene derivatives have shown to be promising anti-TB molecule scaffold due to the low MIC values in acidic pH against MDR M. tuberculosis clinical isolates, synergism with PZA and low cytotoxicity.
Article
Five novel cobalt complexes of pyridine-based heterocyclic thiosemicarbazones [Co(BpT)2]Cl (1), [Co(BpT)2](NO3)•(H2O) (2), [Co(ApT)2]Br•(CH3OH) (3), [Co(BpT)2]Br (4) and [Co(BpT)(OAc)(H2O)2]•2(H2O) (5) have been synthesized (where, HBpT = 2-benzoylpyridine-N⁴-methylthiosemicarbazone, HApT = 2-acetylpyridine-N⁴-phenylthiosemicarbazone) and characterized by analytical and spectroscopic methods. The ligands form monodeprotonated anions in all the complexes to coordinate via thiolate S, azomethine and pyridyl N atoms. The central cobalt metal in the complexes 1, 2, 3 and 4 were found to be in the +3 oxidation state, while in the complex 5 it was found to be in the +2 state. The structures of the complexes 1, 2 and 3 were confirmed by single crystal X-ray diffraction (SCXRD) technique and all of them were crystallized in monoclinic lattices with P21/n space group. The intermolecular interactions were quantified by Hirshfeld surface (HS) analysis. Further, the nature of frontier orbitals and stability of the complexes were theoretically analyzed, and the electrostatic potential plots were mapped on optimized geometries by using computational methods. The complexes 3 and 4 are related by interchanging the phenyl and methyl groups in the thiosemicarbazone moiety. In silico molecular docking study with B-DNA reveals the importance of various interactions and the positioning of the phenyl group. The cobalt(III) complex (3) cation with phenyl group at the N⁴ position is found to have a higher binding tendancy against the therapeutic target B-DNA with docking energy of -9.16 kcal/mol.
Article
The isophthaloylbis(N,N-diphenylthiourea) molecule (H2L ligand) reacts with CuI, CuII, MnIII, FeII, FeIII and CoII ions to form binuclear complexes with formula [M2(L-O,S)2] (M = CuII (1), NiII (2)) or [M2(L-O,S)3] (M = FeIII (4), CoIII (5)). In presence of dimethyl sulfoxide, [Mn2(L-O,S)2(dmso)2] (3) was successfully obtained and represents the first manganese complex with any thiourea derivative ligand. The ligand and the complexes were characterized by spectroscopic techniques and elemental analysis. Complexes 1, 3, 4 and 5 were structurally analyzed by single crystal X-ray diffraction data, confirming the O,S chelating coordination mode. The stoichiometries were confirmed by thermal analysis with formation of the corresponding metal oxides. The set of the performed analysis corroborated that some metal ions undergo oxidation or reduction to form stable 2:2 or 2:3 metal to ligand stoichiometric complexes with H2L.
Article
Full-text available
Herein, we describe the synthesis and characterization of a sulfonate-containing N-donor ligand in its sodium salt and acid forms, Na2[(DES)MeN(CH2)2NMe(DES)]·2H2O (1) and [(DES)MeN+H(CH2)2N+HMe(DES)]·H2O (2), and its corresponding metal complexes, [{(DES)MeN(CH2)2NMe(DES)}M(H2O)2]·nH2O [M = Ni (3), Co (4), Cu (5) and Zn (6)]. Treatment of HIV-infected MT-2 cells with Ni, Co and Cu complexes inhibit virus replication up to 50–70 % both in pre- and post-infected cells as a result of dual preventive and therapeutic behaviour.
Article
Full-text available
Responsible for nearly two million deaths each year, the infectious disease tuberculosis remains a serious global health challenge. The emergence of multidrug- and extensively drug-resistant strains of Mycobacterium tuberculosis confounds control efforts, and new drugs with novel molecular targets are desperately needed. Here we describe lead compounds, the indoleamides, with potent activity against both drug-susceptible and drug-resistant strains of M. tuberculosis by targeting the mycolic acid transporter MmpL3. We identify a single mutation in mmpL3, which confers high resistance to the indoleamide class while remaining susceptible to currently used first- and second-line tuberculosis drugs, indicating a lack of cross-resistance. Importantly, an indoleamide derivative exhibits dose-dependent antimycobacterial activity when orally administered to M. tuberculosis-infected mice. The bioavailability of the indoleamides, combined with their ability to kill tubercle bacilli, indicates great potential for translational developments of this structure class for the treatment of drug-resistant tuberculosis.
Book
This book gives a comprehensive overview about medicinal inorganic chemistry. Topics like targeting strategies, mechanism of action, Pt-based antitumor drugs, radiopharmaceuticals are covered in detail and offer the reader an in-depth overview about this important topic.
Article
Two cobalt(II) complexes [Co(QCT)2]·Cl·1.5H2O (1) (QCT = quinoline-2-carboxaldehyde thiosemicarbazone) and [Co(QCMT)(CH3OH)Cl2] (2) (QCMT = quinoline-2-carboxaldehyde N4-methyl-thiosemicarbazone) have been synthesized and structurally characterized. Complex 1 crystallizes in a triclinic system with space group P–1 and complex 2 crystallizes in a monoclinic system with space group P2(1)/n. In both complexes the cobalt(II) center is six coordinated with distorted octahedral geometry. The interactions of two complexes with CT-DNA were investigated by electronic absorption spectra, circular dichroism (CD) spectra and fluorescence spectra. Results suggest that the complexes bind to DNA via groove binding mode, and complex 2 has stronger binding ability than complex 1. The in vitro cytotoxicity has been tested against the human lung adenocarcinoma cell line A-549, cisplatin-resistant cell line A-549/CDDP, and human breast adenocarcinoma cell line MCF-7. Complex 2 is more cytotoxic than complex 1, and both of them show higher cytotoxicity than the parent ligands alone. Compared with cisplatin, the two cobalt(II) complexes are more active against A-549/CDDP and MCF-7 cell lines at most experimental concentrations. Notably, although complex 2 is found to be less effective than cisplatin against the parent cell line A-549, it is much more effective than cisplatin against the resistant cell A-549/CDDP.
Article
The paper is concerned with the crystal structure and electrochemical characteristics of the cobalt(III) complexes with 2-acetylpyridine S-methylisothiosemicarbazone (HL), of the coordination formulas [CoL2]NO3×MeOH (1), [CoL2]Br×MeOH (2), [CoL2]HSO4×MeOH (3), [CoL2]2[CoII(NCS)4] (4), [Co(HL)(L)]I2×2MeOH (5), and [Co(HL)(L)][CoIICl4]×MeOH (6), as well as of the structure of HL. In all the complexes, Co(III) is situated in a slightly deformed octahedral environment formed by six nitrogen atoms of the two HL molecules in the meridional positions. The ligands coordinate with Co(III) in a usual way, i.e. via the pyridine, azomethine, and isothioamide nitrogen atoms, each forming two five-membered metallocycles. Electrochemical studies of the ligand and the complexes in DMF show that the electrode processes are accompanied by chemical reactions and the corresponding mechanism was proposed. The obtained complexes were also characterized by elemental analysis, conductometric measurements and UV–Vis and IR spectroscopy
Article
Tuberculosis (TB) causes up to 10 million incident cases worldwide per annum. Multidrug-resistant (MDR) and extensively drug-resistant (XDR) strains are leading factors in the resurgence of TB cases and the need to produce new agents to combat such infection. Herein, we describe Co(II) and Cu(II) metal based complexes that feature the pyrophosphate ligand with notable selectivity and marked potency against Mycobacteriumtuberculosis, including MDR strains. Such complexes are confirmed to be bacteriocidal and not affected by efflux inhibitors. Finally, while susceptibility to copper has recently been established for M. tuberculosis, the greater efficacy of cobalt observed herein is of considerable note and in line with the discovery of a copper metallothionein in M. tuberculosis.
Article
Through a systematic variation on the structure of a series of manganese complexes derived from 2-acetylpyridine-N(4)-R-thiosemicarbazones (Hatc-R), structural features have been investigated with the aim of obtaining complexes with potent anti-Mycobacterium tuberculosis activity. The analytical methods used for characterization included FTIR, EPR, UV-visible, elemental analysis, cyclic voltammetry, magnetic susceptibility measurement and single crystal X-ray diffractometry. Density functional theory (DFT) calculations were performed in order to evaluate the contribution of the thiosemicarbazonate ligands on the charge distribution of the complexes by changing the peripheral groups as well as to verify the Mn-donor atoms bond dissociation predisposition. The results obtained are consistent with the monoanionic N,N,S-tridentate coordination of the thiosemicarbazone ligands, resulting in octahedral complexes of the type [Mn(atc-R)2], paramagnetic in the extension of 5 unpaired electrons, whose EPR spectra are consistent for manganese(II). The electrochemical analyses show two nearly reversible processes, which are influenced by the peripheral substituent groups at the N4 position of the atc-R(1-) ligands. The minimal inhibitory concentration (MIC) of these compounds against M. tuberculosis as well as their in vitro cytotoxicity on VERO and J774A.1 cells (IC50) was determined in order to find their selectivity index (SI) (SI=IC50/MIC). The results evidenced that the compounds described here can be considered as promising anti-M. tuberculosis agents, with SI values comparable or better than some commercial drugs available for the tuberculosis treatment.
Article
Transition metal complexes formulated as [Co(L)2]ClO4 (I) and [Ni(L)2] · H2O (II), where HL = pyridine-2-carbaldehyde N(4)-methylthiosemicarbazone, have been synthesized. Complex I was characterized by elemental analysis, IR, MS and single-crystal X-ray diffraction studies. In complex I, the ligand is N2S tridentate, coordinating to the metal center through pyridine nitrogen, imine nitrogen and sulfur atoms. Hydrogen bonds link the different components to stabilize the crystal structure. Preliminary in vitro screening indicated that the free ligand was active against various bacteria and fungi and all the tested compounds showed significant antitumor activity against K562 leukaemia cell line, and can therefore be candidates for further stages of screening in vitro and/or in vivo.