Next Article in Journal
Two Novel Phenolic Compounds from the Rhizomes of Cyperus rotundus L.
Previous Article in Journal
Phenolic Content and Antioxidant Activity of Hibiscus cannabinus L. Seed Extracts after Sequential Solvent Extraction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Improved Antileishmanial Activity of Dppz through Complexation with Antimony(III) and Bismuth(III): Investigation of the Role of the Metal

by
Edgar H. Lizarazo-Jaimes
1,
Rubens L. Monte-Neto
2,
Priscila G. Reis
2,
Nelson G. Fernandes
1,
Nivaldo L. Speziali
3,
Maria N. Melo
4,
Frédéric Frézard
2 and
Cynthia Demicheli
1,*
1
Department of Chemistry, Institute of Exact Sciences, Federal University of Minas Gerais (UFMG), Av. Antônio Carlos 6627, 31270-901, Belo Horizonte, MG, Brazil
2
Department of Physiology and Biophysics, Institute of Biological Sciences, Federal University of Minas Gerais (UFMG), Av. Antônio Carlos 6627, 31270-901, Belo Horizonte, MG, Brazil
3
Department of Physics, Institute of Exact Sciences, Federal University of Minas Gerais (UFMG), Av. Antônio Carlos 6627, 31270-901, Belo Horizonte, MG, Brazil
4
Department of Parasitology, Institute of Biological Sciences, Federal University of Minas Gerais (UFMG), Av. Antônio Carlos 6627, 31270-901, Belo Horizonte, MG, Brazil
*
Author to whom correspondence should be addressed.
Molecules 2012, 17(11), 12622-12635; https://doi.org/10.3390/molecules171112622
Submission received: 24 August 2012 / Revised: 21 October 2012 / Accepted: 22 October 2012 / Published: 25 October 2012
(This article belongs to the Section Medicinal Chemistry)

Abstract

:
Two novel trivalent antimony(III) and bismuth(III) complexes with the nitrogen-donor heterocyclic ligand dipyrido[3,2-a:2',3'-c]phenazine (dppz) were synthesized and characterized as [Sb(dppz)Cl3]∙H2O∙CH3OH and [Bi(dppz)Cl3]. The crystal structure of Sb(III) complex was determined by X-ray crystallography. These complexes were evaluated for their activity against the promastigote form of Sb(III)-sensitive and -resistant Leishmania infantum chagasi and Leishmania amazonensis strains. Both complexes were more effective than dppz alone in inhibiting the growth of Leishmania promastigotes and were at least 77 and 2,400 times more active than potassium antimonyl tartrate in Sb(III)-sensitive and -resistant Leishmania, respectively. The cytotoxicity of dppz and its complexes against mouse peritoneal macrophages occurred at dppz concentrations at least 6-fold greater than those found to be active against Leishmania promastigotes.To investigate the role of the metal in the improved antileishmanial activity of dppz, the activity of the Sb(III) complex was compared between the Sb-resistant mutants and their respective parental sensitive strains. The lack of cross-resistance to the Sb(III)-dppz complex together with the much lower activity of antimonyl tartrate, SbCl3 and BiCl3 strongly support the model that the metal is not active by itself but improves the activity of dppz through complexation.

Graphical Abstract

1. Introduction

Pentavalent antimonials such as sodium stibogluconate and meglumine antimonate, have been used in the treatment of all forms of leishmaniasis for more than half a century [1,2]. Although the mechanism of action of pentavalent antimonial is not fully understood, it is generally accepted that the active form of the metal is the reduced form Sb(III) [3,4].
A major problem in antimonial chemotherapy is the emergence of clinical resistance against pentavalent antimony drugs that has reached epidemic proportions in parts of India [5]. The ATP binding cassette (ABC) protein MRPA plays a major role in metal resistance in Leishmania parasites [6] and its localization in intracellular vesicle membranes suggests that it sequesters Sb(III)-thiol complexes into these vesicles [7].
Other mechanisms such as a diminished biological reduction of Sb(V) to Sb(III) [8], the loss of an aquaglyceroporin (AQP1) allele or its down regulation [9] and hypoxic conditions [10] have been reported to cause an increase in resistance to pentavalent antimonials. In this context, there is a great need for new safe and effective drugs that do not exhibit cross-resistance with conventional antimonial drugs.
Antimony(III) complexes have attracted special interest as potential antineoplastic agents since 1990, when Silvestru et al. [11] reported for the first time the anti-tumor activity of Sb(III) complexes. It was suggested that the mode of action of trivalent antimonial compounds involves some pathways similar to apoptosis, as DNA fragmentation [12,13], which is preceded by an increase in reactive oxygen species caused by alterations of the redox potential [14]. As an antileishmanial agent, Sb(III) inhibits and forms a complex with the enzyme trypanothione reductase which acts by recycling trypanotione disulphides in trypanothione, the major antioxidant thiol in Leishmania, important to maintain their intracellular redox balance [15,16].
On the other hand, bismuth compounds have been used to treat infections caused by Helicobacter Pylori bacteria and duodenal ulcers [17], as well as radio-therapeutic agents for cancer treatment [18]. Despite the close periodic table relationship of antimony and bismuth, there are only few reports in the literature of bismuth-based drugs being developed and evaluated as antileishmanial agents [19]. On the other hand, new bismuth complexes with DNA affinity and activity against cancer cell were reported [20,21].
Metal complexes with polypyridyl ligands, have been extensively studied by a number of research groups over the last year due to potential application in photodynamic therapy [22,23] and probes for biological molecules [24,25].
Recently, complexes of dipyrido[3,2-a:2',3'-c]phenazine (dppz) with gold (Au), copper (Cu) and vanadium (V) have shown remarkable antileishmanial or antitrypanosomal activities [26,27], however, the possible synergism between the metal and the ligand was not fully investigated. On the other hand, there are few reports in the literature referring to the structural characterization of Sb(III) and Bi(III) complexes with polypyridyl ligands [28,29], but no report on their pharmacological activities.
In this work, the synthesis, the structural and physicochemical characterization of the polypyridyl ligand dppz coordinated to Sb(III) and Bi(III) are reported. The activities of dppz and its resulting complexes against both Sb(III)-sensitive and -resistant Leishmania infantum chagasi and Leishmania amazonensis promastigotes and their cytotoxicities towards peritoneal macrophages are described. The present paper reports for the first time a new experimental approach, based on tests against metal-resistant mutants and their respective parental sensitive strains, to get insight into the role of the metal in the cytotoxicity of a metal complex.

2. Results and Discussion

Antimony and bismuth complexes [Sb(dppz)Cl3] (1) and [Bi(dppz)Cl3] (2) were obtained in good yields by refluxing equimolar amounts of SbCl3 or BiCl3 with dppz in methanolic solutions. The complexes 1 and 2 prepared during the course of this investigation are an air-stable crystal and powder, respectively, which were characterized by spectroscopic methods including FT-IR, NMR, elemental analysis and also (in the case of complex 1) by X-ray crystallography. In the case of complex 2, the crystals obtained did not show sufficient quality for X-ray analysis. In FT-IR spectra of complexes 1 and 2, bands were observed for C=N and C=C bonds stretching in aromatics in the 1640-1400 cm−1 range. The dppz bands at around 1632 and 1616 cm−1 are shifted considerably towards lower frequency—1616 and 1602 cm−1 for [Sb(dppz)Cl3] and 1616 and 1598 cm−1 for [Bi(dppz)Cl3]—confirming the coordination of N atoms to Sb and Bi. Similar results for the spectroscopic data of dppz complexes with Au and Cu have been reported previously [26]. The corresponding 1H-NMR spectra showed only one set of signals, shifted downfield in both complexes with respect to the free ligand (Table 1). 1H-NMR spectra of [Sb(dppz)Cl3] and [Bi(dppz)Cl3] complexes in DMSO-d6 exhibited the presence of the polypyridine ligand aromatic protons between δ 9.68 and 8.05 ppm.
Table 1. Selected 1H-NMR and IR data for free ligand and complexes 1 and 2.
Table 1. Selected 1H-NMR and IR data for free ligand and complexes 1 and 2.
Compound1H-NMR data (ppm)References
HcHaHdHeHb
Dppz9.559.238.418.077.95Navarro et al. (2006) [26]
Sb(dppz)Cl39.689.318.398.188.11This work
Bi(dppz)Cl39.549.398.328.078.05
The Hc proton of dppz and both complexes showed resonances at very low field because they are also highly deshielded due to the anisotropic effect provided by the nitrogen atom of the pyrazine ring (Figure 1). Ha protons in proximity of the coordinating nitrogen atoms of dppz ligand experienced a downfield shift by about δ 0.08-0.16 in complexes 1 and 2, respectively, indicating the involvement of the dppz nitrogens in coordination. The unequivocal proton NMR assignments were made by concerted analysis of 1D 1H-, 1D 13C-, 2D 1H 1H COSY and 2D 1H 13C-HMQC.
Figure 1. Molecular structure of [M(dppz)Cl3], with M= Sb or Bi.
Figure 1. Molecular structure of [M(dppz)Cl3], with M= Sb or Bi.
Molecules 17 12622 g001
For both complexes, satisfactory elemental analyses of C, H and N were obtained. In complex 1 crystallization methanol and water molecules are present, as confirmed by its thermogravimetric curves which showed a weight loss of 8.26% (calc. 8.92%), nevertheless complex 2 did not have crystallization molecules. The asymmetric unit of complex 1 is shown in Figure 2. Crystal data and structure refinement parameter are given in Table 2 and the selected bonds length and angles are given in Table 3.
Figure 2. Molecular structure of [Sb(dppz)Cl3].
Figure 2. Molecular structure of [Sb(dppz)Cl3].
Molecules 17 12622 g002
Table 2. Crystal data and structure refinement parameter for complex 1.
Table 2. Crystal data and structure refinement parameter for complex 1.
Sb(dppz)Cl3
Empirical formulaC19H10Cl3 N4O1.25Sb
Formula weight 542.41
Temperature (K)293(2)
Crystal systemMonoclinic
Space groupP2(1)/c
a (Å) 10.3244(2)
b (Å)13.2760(3)
c (Å)14.5753(3)
α (o)90
β (o)92.0930(19)
γ (o)90
V3)1994.44(7)
Z4
F (000)1056
Dcalc (mg/m3)1.805
Crystal dimensions mm30.2 × 0.2 × 0.3
θ Range (o)4.29-62.65
Reflections collected9035
Independent reflection3138
Rint0.0355
Maximum/minimum transmission1.000 / 0.30303
Data/restraints/parameters3138 / 0 / 247
Goodness-of-fit on F21.060
Final R indices R1 = 0.0352, wR2 = 0.0826
[I > 2σ(I)]2478
R indices (all data)R1 = 0.0522, wR2 = 0.0886
Largest difference in peak/hole (e·A−3)0.761 and −0.560
Table 3. Selected bond length (Å) and angles (°) for complex 1.
Table 3. Selected bond length (Å) and angles (°) for complex 1.
Sb-N(2)2.245(4)N(2)-Sb-N(1)71.44(14)
Sb-N(1)2.345(4)N(1)-Sb-Cl(3)159.07(10)
Sb-Cl(3)2.4992(15)N(2)-Sb-Cl(3)87.73(11)
Sb-Cl(1)2.5126(13)N(2)-Sb-Cl(1)82.14(10)
Sb-Cl(2)2.6348(14)N(1)-Sb-Cl(1)84.23(10)
N(1)-C(16)1.334(6)Cl(3)-Sb-Cl(1)95.22(5)
N(1)-C(12)1.351(6)N(2)-Sb-Cl(2)80.25(10)
N(2)-C(22)1.336(6)N(1)-Sb-Cl(2)82.97(10)
N(2)-C(26)1.368(6)Cl(3)-Sb-Cl(2)91.68(5)
Cl(1)-Sb-Cl(2)160.80(5)
The geometry around the metal center is distorted square pyramidal (SP), with N(2) as the apex and Cl(1), Cl(2), Cl(3), N(1) occupying the equatorial plane. The four equatorial bonds are long (means 2,498 Å), while the apical donor atom form the strongest bond (Sb(1)-N(2) 2.245(4) Å). The sum of equatorial angles is 354° (Cl(3)-Sb-Cl(1) 95.22(5)°, Cl(3)-Sb-Cl(2) 91.68(5)°, N(1)-Sb(1)-Cl(1) 84.23(10)°, N(1)-Sb(1)-Cl(2) 82.97(10)°), which suggest some distortion from an ideal square pyramidal [30]. As shown in Figure 2, the asymmetrical unit in the crystal contains one molecule of methanol and one of water besides the [Sb(dppz)Cl3] molecule.
Both complexes 1 and 2 were found to be highly active against Leishmania promastigote forms (Figure 3). The IC50 for complexes 1 and 2 in the wild-type (WT) strains were similar, corresponding to ~0.6 and ~1 μM for L. infantum chagasi and L. amazonensis, respectively (Table 4).
Strikingly, the IC50 values for dppz complexes were similar when comparing the Sb-resistant mutants with their respective parental sensitive WT strains (Table 4). This is in contrast with the trivalent potassium antimonyl tartrate that was much less active against Sb-resistant mutants (Table 4).
Figure 3. Dppz-derived Sb(III) and Bi(III) complexes sensitivity profiles of antimony-sensitive and -resistant L. chagasi and L. amazonensis promastigotes. (A) L. chagasi BH400 wild-type; (B) L. amazonensis BA199 wild-type; (C) L. chagasi BH400Sb2700.2 (Sb-resistant); (D) L. amazonensis BA199Sb2700.2 (Sb-resistant). (•) dppz; (□) dppz/Bi; (∆) dppz/Sb. The Leishmania promastigotes were cultivated in 24-well cell culture plates containing 2 mL of α-MEM medium for 72 h at 25 °C. The resistant lines were previously selected in vitro by step wise drug selection. All data represent average of, at least, three independent experiments.
Figure 3. Dppz-derived Sb(III) and Bi(III) complexes sensitivity profiles of antimony-sensitive and -resistant L. chagasi and L. amazonensis promastigotes. (A) L. chagasi BH400 wild-type; (B) L. amazonensis BA199 wild-type; (C) L. chagasi BH400Sb2700.2 (Sb-resistant); (D) L. amazonensis BA199Sb2700.2 (Sb-resistant). (•) dppz; (□) dppz/Bi; (∆) dppz/Sb. The Leishmania promastigotes were cultivated in 24-well cell culture plates containing 2 mL of α-MEM medium for 72 h at 25 °C. The resistant lines were previously selected in vitro by step wise drug selection. All data represent average of, at least, three independent experiments.
Molecules 17 12622 g003
Table 4. Inhibitory concentrations of dppz and its Sb(III) and Bi(III) complexes towards antimony-sensitive and -resistant New World Leishmania species.
Table 4. Inhibitory concentrations of dppz and its Sb(III) and Bi(III) complexes towards antimony-sensitive and -resistant New World Leishmania species.
StrainIC50 (μM) ± SEM (CI 95%)
DppzdppzBiCl3dppzSbCl3TA *SbCl3BiCl3
L. infantum chagasi (WT)0.81 ± 0.040.59 ± 0.010.62 ± 0.01100 ± 3341 ± 2563 ± 2
(0.71–0.9)(0.56–0.63)(0.59–0.65)
L. infantum chagasi (SbR)1.86 ± 0.080.61 ± 0.020.57 ± 0.01>2,700462 ± 3>500
(1.69–2.03)(0.56–0.66)(0.55–0.60)
L. amazonensis (WT)>21.07 ± 0.030.95 ± 0.0283 ± 1362 ± 2621 ± 2
(1.01–1.14)(0.89–1.01)
L. amazonensis (SbR)2.00 ± 0.051.12 ± 0.010.92 ± 0.02>2,7001503 ± 2>500
(1.89–2.10)(1.08–1.16)(0.86-0.97)
* TA: potassium antimonyl tartrate as Sb(III) source; SEM:standard error of the mean; CI: confidence interval; SbR (Sb-resistant); The IC50 values were calculated by non-linear regression.
When tested alone, dppz showed intrinsic activities against Leishmania, with IC50 varying from 0.8 to ~2 μM for both Sb-sensitive or -resistant mutants of the two species studied (Table 4). Interestingly, complexes 1 and 2exhibited significantly higher antileishmanial activity than dppz itself, indicating an additional role of the metal (Sb(III) and Bi(III)) in the antileishmanial activity of the complexes (Figure 3 and Table 4). Since complexes 1 and 2 were at least 80-fold more active against WT Leishmania strains than trivalent potassium antimonyl tartrate, BiCl3 and SbCl3 (Table 4), one can infer that the metal may not be active by itself and that the higher activity of the complexes compared to dppz may be related to the improved activity of dppz through metal complexation. The lower IC50 value of antimonyl tartrate compared to SbCl3 in WT L. amazonensis also illustrates that complexation of a cytotoxic metal ion may enhance its activity, presumably through improved delivery to the target. In the case of antimonyl tartrate, much greater IC50 values were found in Sb-resistant mutants compared to the respective WT parental cells, demonstrating clearly that the metal ion contributes to the cytotoxicity of this Sb(III) complex. This is in contrast with the Sb(III)-dppz complex that was equally active against Sb-resistant and -sensitive Leishmania strains. Thus, the latter observation further supports the model that the metal is not directly responsible for the antileshmanial activity of Sb(III)-dppz complex. As shown in Table 5, the cytotoxicity of dppz and its complexes against mouse peritoneal macrophages occurred at dppz concentrations at least 6-fold greater than those found to be active against Leishmania promastigotes, indicating the selectivity of the complexes towards Leishmania parasites.
Table 5. Cytotoxicity of dppz and its Sb(III) and Bi(III) complexes against mouse peritoneal macrophages.
Table 5. Cytotoxicity of dppz and its Sb(III) and Bi(III) complexes against mouse peritoneal macrophages.
Dppz[Bi(dppz)Cl3][Sb(dppz)Cl3]
CC50 (μM) *12.54.87.0
SI #15.48.111.3
* CC50: concentration of dppz which is cytotoxic against 50% of macrophages; # SI: selective index, calculated as the ratio between CC50 in murine macrophages and IC50 in L. infantum chagasi (WT).
The L. infantum chagasi is the etiological agent for visceral leishmaniasis [31] while L. amazonensis is related to the cutaneous form of the disease in the New World [32]. Since these two species are causative agents of different clinical manifestations, they might also differ metabolically which reflect in their distinct drug sensitivity. This observation highlights the importance to use different Leishmania species for drug prospecting screening [33]. In addition to the high antileishmanial activity of 1 and 2, both complexes can bypass the Sb-resistance of Leishmania in two different species of Sb-resistant Leishmania mutants (Table 4). The Sb-resistant selected lines of L. infantum chagasi BH400 Sb2700.2 and L. amazonensis BA199 Sb2700.2 were able to grow in the presence of Sb concentrations greater than 2700 μM, but failed to grow up in the presence of ≤1.12 μM of 1 and/or 2 (Table 4). Thus, this data taken altogether indicates that complexes 1 and 2 are very promising drug candidates for the treatment of Sb-resistant leishmaniasis.
An important physical characteristic that affects the bioavailability of a compound is its lipophilicity. Therefore a relatively optimal lipophilicity should be critical for a good drug candidate.
The lipophilicity of the complexes 1 and 2 was determined by measuring their octanol/water partition coefficients (log P). The log P values of the free dppz ligand was calculated using the online program ALOGSP 2.1. The log P values for dppz and complexes 1 and 2 are respectively 2.77 ± 0.76; 0.94 ± 0.24 and 0.98 ± 0.17. More positive log P values correspond to more lipophilic compounds, whereas more negative log P values correspond to more hydrophilic ones. This data indicates that complexation with Sb(III) and Bi(III) makes dppz less lipophilic. Interestingly, the antileishmanial activity profile follows the hydrophilicity profile of the compound, i.e., the more hydrophilic, the more active.
A possible explanation for the leishmanicidal activity of dppz complexes is their interaction with parasite DNA through intercalation, a mechanism recently suggested by Navarro and co-workers [34]. They described both Cu(II) and Au(III) complexes with dppz showing a potent leishmanicidal activity against both promastigotes of Leishmania (Leishmania) mexicana and Leishmania (Viannia) braziliensis [34]. Although the best-known type of interaction of the polypyridyl complexes with DNA is the intercalation, highly reactive metal complexes containing reactive Cl could form the covalent bonds with DNA, as reported for [Ru(tpy)(pap)(CH3CN)](ClO4)2and [Ru(tpy)(dppz)(CH3CN)]2+[35,36]. In future studies, the interaction of Sb(III) and Bi(III) dppz complexes with DNA should be investigated, as a possible mechanism of cytotoxicity. In our work the differences in observed cytotoxicities of the complexes compared to dppz could be attributed to the changes in dppz lipophilicity following the coordination by SbCl3 and BiCl3. The metal complexation to dppz probably leads to an increase of dppz bioavailability.

3. Experimental

3.1. General

Commercially available methanol was dried before use. All the preparations were done under anhydrous conditions. dppz was synthesized by a condensation of 1,10-phenanthroline-5,6-dione and ortho-phenylenediamine according to the procedure described by Liang [37]. Sb(III) chloride (SbCl3) and Bi(III) chloride (BiCl3), from Sigma-Aldrich (St. Louis, MO, USA) were used without further purification. Elemental analyses for C, H and N were carried out with a Perkin Elmer PE-2400 instrument. IR spectra in the 4,000–400 cm−1 region were obtained for KBr pellets, with a Perkin Elmer RX-83303. The absorption spectra were recorded on a Varian Cary 100 UV-Visible Spectrophotometer.
NMR spectra were recorded at 400 MHz using Bruker DPX-400 spectrometer. 1H- and 13C-NMR chemical shifts were measured relative to tetramethylsilane (TMS) and dimethyl sulfoxide-d6 (DMSO-d6) as the solvent. Intensity data for the X-ray were collected with Xcalibru Atlas Gemini ultra. Crystallographic data and experimental details of the structure determinations are listed in Table 1.

3.2. Synthesis of the Complexes [Sb(dppz)Cl3] and [Bi(dppz)Cl3]

3.2.1. Synthesis of [Sb(dppz)Cl3] (1)

A solution of dipyrido[3,2-a:2',3'-c]phenazine (0.05 g, 0.18 mmol) in dry methanol (26 mL) was added slowly to a freshly prepared antimony trichloride (0.041 g SbCl3, 0.18 mmol in 10 mL of dry methanol) solution at room temperature. The reaction mixture was stirred for 6 hours at 40 °C. The resulting clear solution was kept in the darkness at room temperature to give yellow crystals. Yield: 82.9% (0.92 g). IR (KBr, cm−1): 1614, 1578 ν(C=C, C=N); 1,138, 818, 724 δ(Csp2-H). 1H-NMR (DMSO-d6 400 MHz): δ 9.68 (dd, Hc); 9.31 (dd, Ha); 8.39 (q, Hd); 8.21–8.08 (m, He and Hb). Thermogravimetric and crystal data indicated the presence of one crystallization molecule of water and methanol. Elemental analysis for C19H16Cl3N4O2Sb, calc. (found): C% 40.70 (41.21); H% 2.86 (2.73); N% 10.00 (10.25).

3.2.2. Synthesis of [Bi(dppz)Cl3] (2)

A solution of dipyrido[3,2-a:2',3'-c]phenazine (0.07 g, 0.25 mmol) in dry methanol (26 mL) was added slowly to a freshly prepared antimony trichloride (0.079 g BiCl3, 0.25 mmol in 10 mL of dry methanol) solution at room temperature. The reaction mixture was stirred for 3 h at 40 °C and the resulting precipitatate was filtered and dried. Yield: 82.9% (0.123 g). IR (KBr, cm−1): 1,614, 1,574 ν(C=C, C=N); 1,140, 816, 730 δ(Csp2-H). 1H-NMR (DMSO-d6 400 MHz): δ 9.54 (dd, Hc); 9.39 (dd, Ha); 8.32 (q, Hd); 8.04–8.08 (m, He and Hb). Elemental analysis for C18H12Cl3N4Bi, calc. (found): C% 36.15 (36.21); H% 1.67 (1.63); N% 9.37 (9.19).

3.2.3. Stability of the Complexes

The stability of the complexes was checked by UV absorption spectroscopy. For this purpose, UV absorption spectra were registered at different time intervals (from 0 to 4 h) for dppz and complexes 1 and 2 at 10 µmol·L−1 in DMSO, DMSO-containing water and octanol. The spectra of the complexes were different from that of dppz and did not show significant changes as a function of time.

3.3. X-ray Crystallography

Single crystal X-ray diffraction data were obtained at room temperature on a Xcalibur Atlas Gemini Ultra diffractometer, using graphite monochromated MoKα radiation (λ = 0.71069 Å). Final unit cell parameters and the integration of the collected reflections were performed using the CRYSALISPRO software (Version 1.171.33.55 release 05-01-2010 CrysAlis171.NET). The structure solutions and full-matrix least-squares refinements based on F2 were performed with the SHELXS-97 and SHELXL-97 program package [38]. All atoms except hydrogen were refined anisotropically. Although many hydrogen atoms could be identified in a Fourier difference map all of them were geometrically added to the structure and then refined by the riding model in the final stages. Details of data collection and structure refinement are given in Table 2. Selected distances and angles are given in Table 3. The crystallographic data were deposited at Cambridge Crystallographic Data Center on CCDC 874746.

3.4. Antileishmanial and Cytotoxic Activities

3.4.1. Parasite Culture

Leishmania (Leishmania) amazonensis (strain MHOM/BR/1989/BA199) and Leishmania (Leishmania)infantum chagasi (strain MCAN/BR/2002/BH400) promastigotes were maintained in minimum essential culture medium (α-MEM) (Gibco, Invitrogen, Grand Island, NY, USA) supplemented with 10% (v/v) heat inactivated fetal calf serum (Cultilab, Campinas, SP, Brazil), 100 mg/mL kanamycin, 50 mg/mL ampicillin, 2 mM L-glutamine, 5 mg/mL hemin, 5 mM biopterin, (Sigma-Aldrich), pH 7.0 and incubated at 25 °C. L. amazonensis and L. infantum chagasi were selected for Sb(III) resistance as previously decribed [39,40]. The Sb(III)-resistant mutants L. amazonensis BA199Sb2700.2 and L. infantum chagasi BH400Sb2700.2 were selected in 25 cm2 flasks containing 5 mL of α-MEM in the presence of increasing Sb(III) concentrations up to 2,700 μM.

3.4.2. Antileishmanial Activity Assay

Complexes 1 and 2 were evaluated in vitro for their activity against both Sb(III)-sensitive and -resistant Leishmania parasites. Log-phase L. amazonensis and L. infantum chagasi promastigotes (1 × 106 parasites/mL) were seeded in 24-wells cell culture plates with 1.5 mL of α-MEM, incubated under shaking at 25 °C during 72 h in the presence of several concentrations of complex 1 or 2. Controls were performed using cultures in the presence of potassium antimonyl tartrate, SbCl3, BiCl3 and dppz itself. Non-treated parasites were established for growth comparison. Stock solutions of dppz, its complexes, SbCl3 and BiCl3 were dissolved in DMSO (5 M) and diluted in α-MEM cell culture medium to obtain the range of tested concentrations. The final DMSO concentration did not exceed 0.2%, which is known to be nontoxic to Leishmania parasites [41,42]. For drug susceptibility assay, Leishmania growth curves were constructed by measuring absorbance at 600 nm [43]. The antileishmanial activity is expressed as IC50/72 h, which is the concentration that reduces cell growth by 50% compared to untreated control (relative growth). All experiments were done at least three times as independent experiments performed in triplicate.

3.4.3. Cytotoxicity Assay against Peritoneal Macrophages

The concentration of studied compounds which is cytotoxic to 50% of the macrophages (CC50) was determined by 3-(4,5-dimethylthiazol-2-yl)-2,5 diphenyltetrazolium bromide (MTT) method according to Mosmann [44]. Macrophages were obtained by lavage of peritoneal cavity of mice with 10 mL cold RPMI without FBS. After washing, the cell suspension (4.0 × 106/mL) was seeded (0.1 mL) in 96-well flat bottom plates. Macrophages were allowed to adhere for 2 h and non-adherent cells were removed by washing with RPMI. Then, the studied compounds were added to the wells at concentrations ranging from 1 to 20 μM and the cells were further cultured in RPMI-1640 supplemented with 10% FBS for 24 h at 37 °C in a humidified 5% CO2 atmosphere. Thereafter, the medium was replaced with fresh RPMI containing 0.5 mg/mL of MTT and the plates were incubated for additional 4 h. Supernatants were aspirated and the formazan crystals formed were dissolved in 100 μL of DMSO. After 15 min of incubation at room temperature, absorbance of solubilized MTT formazan product was spectrophotometrically measured at 570 nm.

3.5. Statistical Analysis

The IC50 and CC50 values were calculated by nonlinear regression using the software GraphPad Prism 5.0. The acceptable level of significance was 95% (p < 0.05).

3.6. n-Octanol/Water Partition Coefficient

Log P values were determined by the Shake-Flask method dissolving complexes 1 and 2 in octanol-saturated aqueous phosphate buffer (pH 6.8) to a concentration at which the solution was still unsaturated but close to saturation, 10 and 5 µmol·L−1, respectively. Then, three different partition system were prepared by adding these solutions to water-saturated octanol, in octanol/water volume ratios, 1:1, 1:2 and 2:1 in duplicate. The mixtures were vortexed for 10 min and then centrifuged for 5 min. The partition coefficients were determined using the absorbance of complexes 1 and 2. Calculated dppz log P value was obtained from the program ALOGPS 2.143 available on line. The feasibility of ALOGPS 2.143 ability to predict partition coefficient was successfully demonstrated [45].

4. Conclusions

Two novel complexes of Sb(III) and Bi(III) with dppz were synthesized in reasonably good yields and characterized as [Sb(dppz)Cl3]∙H2O∙CH3OH and [Bi(dppz)Cl3]. These complexes were more effective than dppz alone and at least 77 and 2,400 times more active than potassium antimonyl tartrate against Sb(III)-sensitive and -resistant Leishmania, respectively. The lack of cross-resistance to the Sb(III)-dppz complex, together with the much lower activity of antimonyl tartrate, SbCl3 and BiCl3 strongly support the model that the metal is not active by itself but improves the activity of dppz through complexation. Complexation was found to decrease the lipophilicity of dppz, suggesting the role of this factor in the improved antileishmanial activity of the complexes. The present work successfully applied a novel approach, based on tests against metal-resistant mutants and their respective parental sensitive strains, to get insight into the role of the metal in the cytotoxicity of metal complexes.

Acknowledgments

The authors are grateful to the Brazilian agencies CNPq, FAPEMIG and CAPES for financial support.

References

  1. Marsden, P.D. Pentavalent antimonials: Old drugs for new diseases. Rev. Soc. Bras. Med. Trop. 1985, 18, 187–198. [Google Scholar]
  2. Frézard, F.; Demicheli, C.; Ribeiro, R.R. Pentavalent antimonials: New perspectives for old drugs. Molecules 2009, 14, 2317–2336. [Google Scholar]
  3. Frézard, F.; Demicheli, C.; Ferreira, C.S.; Costa, M.A.P. Glutathione-induced conversion of pentavalent antimony to trivalent antimony in meglumine antimoniate. Antimicrob. Agents Chemother. 2001, 45, 913–916. [Google Scholar]
  4. Ferreira, C.D.; Martins, P.S.; Demicheli, C.; Brochu, C.; Ouellette, M.; Frézard, F. Thiol-induced reduction of antimony(V) into antimony(III): A comparative study with trypanothione, cysteinyl-glycine, cysteine and glutathion. Biometals 2003, 16, 441–446. [Google Scholar] [CrossRef]
  5. Sundar, S.; More, D.K.; Singh, M.K.; Singh, V.P.; Sharma, S.; Makharia, A.; Kumar, P.C.K.; Murray, H.W. Failure of pentavalent antimony in visceral leishmaniasis in India: Report from the center of the Indian epidemic. Clin. Infect. Dis. 2000, 31, 1104–1107. [Google Scholar]
  6. Ouellette, M.; Légaré, D.; Haimeur, A.; Grondin, K.; Roy, G.; Brochu, C.; Papadopoulou, B. ABC transporters in Leishmania and their role in drug resistance. Drug Resist. Updat. 1998, 1, 43–48. [Google Scholar] [CrossRef]
  7. Légaré, D.; Richard, D.; Mukhopadhyay, R.; Stierhof, Y.D.; Rosen, B.P.; Haimeur, A.; Papadopoulou, B.; Ouellette, M. The Leishmania ATP-binding cassette protein PGPA is an intracellular metal-thiol transporter ATPase. J. Biol. Chem. 2001, 276, 26301–26307. [Google Scholar]
  8. Shaked-Mishan, P.; Ulrich, N.; Ephros, M.; Zilberstein, D. Novel intracellular SbV reducing activity correlates with antimony susceptibility in Leishmania donovani. J. Biol. Chem. 2001, 276, 3971–3976. [Google Scholar]
  9. Gourbal, B.; Sonuc, N.; Bhattacharjee, H.; Légaré, D.; Sundar, S.; Ouellette, M.; Rosen, B.P.; Mukhopadhyay, R. Drug uptake and modulation of drug resistance in Leishmania by an aquaglyceroporin. J. Biol. Chem. 2004, 279, 31010–31017. [Google Scholar]
  10. Ayres, D.C.; Pinto, L.A.; Giorgio, S. Efficacy of pentavalent antimony, amphotericin B, and miltefosine in Leishmania amazonensis-infected macrophages under normoxic and hypoxic conditions. J. Parasitol. 2008, 94, 1415–1417. [Google Scholar] [CrossRef]
  11. Silvestru, C.; Socaciu, C.; Bara, A.; Haiduc, I. The first organoantimony(III) compounds possessing antitumor properties: Diphenylantimony(III) derivatives of dithiophosphorus ligand. Anticancer Res. 1990, 10, 803–804. [Google Scholar]
  12. Sereno, D.; Holzmuller, P.; Mangot, I.; Cuny, G.; Ouaissi, A.; Lemesre, J.L. Antimonial-mediated DNA fragmentation in Leishmania infantum amastigotes. Antimicrob. Agents Chemother. 2001, 45, 2064–2069. [Google Scholar] [CrossRef]
  13. Moreira, W.; Leprohon, P.; Ouellette, M. Tolerance to drug-induced cell death favours the acquisition of multidrug resistance in Leishmania. Cell Death Dis. 2011, 2, 1–8. [Google Scholar]
  14. Wyllie, S.; Fairlamb, A.H. Differential toxicity of antimonial compounds and their effects on glutathione homeostasis in a human leukaemia monocyte cell line. Biochem. Pharmacol. 2006, 71, 257–267. [Google Scholar]
  15. Cunningham, M.L.; Fairlamb, A.H. Trypanothione reductase from Leishmania donovani Purification, characterization and inhibition by trivalent antimonials. Eur. J. Biochem. 1995, 230, 460–468. [Google Scholar] [CrossRef]
  16. Baiocco, P.; Colotti, G.; Franceschini, S.; Ilari, A. Molecular basis of antimony treatment in leishmaniasis. J. Med. Chem. 2009, 52, 2603–2612. [Google Scholar]
  17. Klapotke, T. Biological activity of organometallic bismuth compounds. Biol. Met. 1988, 1, 69–76. [Google Scholar]
  18. Yang, N.; Sun, H. Biocoordination chemistry of bismuth: Recent advances Coord. Chem. Rev. 2007, 251, 2354–2366. [Google Scholar]
  19. Andrews, P.C.; Frank, R.; Junk, P.C.; Kedzierski, L.; Kumar, I.; MacLellan, J.G. Anti-Leishmanial activity of homo- and heteroleptic bismuth(III) carboxylates. J. Inorg. Biochem. 2011, 105, 454–461. [Google Scholar]
  20. Ensafi, A.A.; Hajian, R.; Ebrahimi, S. Study on the interaction between Morin-Bi(III) complex and DNA with the use of methylene blue dye as a fluorophor probe. J. Braz. Chem. Soc. 2009, 20, 266–276. [Google Scholar] [CrossRef]
  21. Oliveira, L.G.; Silva, M.M.; de Paula, F.C.S.; Pereira-Maia, E.C.; Donnici, C.L.; de Simone, C.A.; Frézard, F.; da Silva Junior, E.N.; Demicheli, C. Antimony(V) and bismuth(V) complexes of lapachol: Synthesis, crystal structure and cytotoxic activity. Molecules 2011, 16, 10314–10323. [Google Scholar] [CrossRef]
  22. Sasmal, P.K.; Patra, A.K.; Nethaji, M.; Chakravarty, A.R. DNA cleavage by new oxovanadium(IV) complexes of N-salicylidene α-amino acids and phenanthroline bases in the photodynamic therapy window. Inorg. Chem. 2007, 46, 11112–11121. [Google Scholar]
  23. Toneatto, J.; Boero, R.A.; Lorenzatti, G.; Cabanillas, A.M.; Argueello, G.A. New insights in the DNA-[Cr(phen)2(dppz)]3+ binding and photocleavage properties by the complex with an intercalating ligand. J. Inorg. Biochem. 2010, 104, 697–703. [Google Scholar]
  24. Friedman, A.E.; Chambron, J.C.; Sauvage, J.P.; Turro, N.J.; Barton, J.K. Molecular light switch for DNA: Ru(bpy)2(dppz)2+. J. Am. Chem. Soc. 1990, 112, 4960–4962. [Google Scholar] [CrossRef]
  25. Zhang, K.Y.; Li, S.P.Y.; Zhu, N.; Or, I.W.S.; Cheung, M.S.H.; Lam, Y.W.; Lo, K.K.W. Structure, photophysical and electrochemical properties, biomolecular interactions, and intracellular uptake of luminescent cyclometalated iridium(III) dipyridoquinoxaline complexes. Inorg. Chem. 2010, 49, 2530–2540. [Google Scholar] [CrossRef]
  26. Navarro, M.; Hernandez, C.; Colmenares, I.; Hernandez, P.; Fernandez, M.; Sierralta, A.; Marchan, E. Synthesis and characterization of [Au(dppz)2]Cl3: DNA Interaction studies and biological activity against Leishmania (L) mexicana. J. Inorg. Biochem. 2007, 101, 111–116. [Google Scholar] [CrossRef]
  27. Benítez, J.; Guggeri, L.; Tomaz, I.; Pessoa, J.C.; Moreno, V.; Lorenzo, J.; Avilés, F.X.; Garat, B.; Gambino, D. A novel vanadyl complex with a polypyridyl DNA intercalator as ligand: A potential anti-protozoa and anti-tumor agent. J. Inorg. Biochem. 2009, 103, 1386–1387. [Google Scholar]
  28. Yin, H.D.; Zhai, J. Synthesis, characterizations and crystal structures of antimony(III) complexes with nitrogen-containing ligands. Inorg. Chim. Acta 2009, 362, 339–345. [Google Scholar] [CrossRef]
  29. Espínola, J.G.P.; Martins, E.P.S.; Aguiar, F.P.; Silva, H.R.M.; Fonseca, M.G.; Arakaki, L.N.H.; Teotôno, E.E.S. Thermal decomposition study of bismuth (III) trichloride complex with 1,10-phenanthroline as the ligand. J. Therm. Anal. Calorim. 2011, 106, 601–606. [Google Scholar]
  30. Bertazzi, N.; Alonzo, G.; Gibb, T.C. Antimony-121 moessbauer and infrared spectral studies on 1,10-phenanthroline adducts of antimony(III) halides. Inorg. Chim. Acta 1983, 73, 121–124. [Google Scholar] [CrossRef]
  31. Shaw, J.J. Further thoughts on the use of the name Leishmania (Leishmania) infantum chagasi for the a etiological agent of American visceral leishmaniasis. Mem. Inst. Oswaldo Cruz 2006, 101, 577–579. [Google Scholar] [CrossRef]
  32. Marzochi, M.C.D.A.; Marzochi, K.B.F. Tegumentary and visceral leishmaniasis in Brazil emerging anthropozoonosis and possibilities for their control. Cad. Saúde Públ. 1994, 10, 359–375. [Google Scholar] [CrossRef]
  33. Sinagra, A.; Luna, C.; Abraham, D.; Iannella, M.D.C.; Riarte, A.; Krolewiecki, A.J. The activity of azithromycin against Leishmania (Viannia) braziliensis and Leishmania (Leishmania) amazonensis in the golden hamster model. Rev. Soc. Bras. Med. Trop. 2007, 40, 627–630. [Google Scholar] [CrossRef]
  34. Navarro, M.; Cisneros-Fajardo, E.J.; Sierralta, A.; Fernandez-Mestre, M.; Silva, P.; Arrieche, D.; Marchán, E. Design of copper DNA intercalators with leishmanicidal activity. J. Biol. Inorg. Chem. 2003, 8, 401–408. [Google Scholar]
  35. Pramanik, N.C.; Pramanik, K.; Ghosh, P.; Bhattacharya, S. Chemistry of [Ru(tpy)(pap)(L')n+ (tpy = 2,2',6',2 ''-terpyridine; pap = 2-(phenylazo)pyridine; L' = Cl, H2O, CH3CN, 4-picoline, N3; n = 1,2. X-ray crystal structure of [Ru(tpy)(pap)(CH3CN)](ClO4)2 and catalytic oxidation of water to dioxygen by [Ru(tpy)(pap)(H2O)]2+. Polyhedron 1998, 17, 1525–1534. [Google Scholar] [CrossRef]
  36. Ossipov, D.; Gohil, S.; Chattopadhyaya, J. Synthesis of the DNA-[Ru(tpy)(dppz)(CH3CN)]2+ conjugates and their photo cross-linking studies with the complementary DNA strand. J. Am. Chem. Soc. 2002, 124, 13416–13433. [Google Scholar]
  37. Liang, B.; Zhu, M.X.; Zhu, W.G. Synthesis and Photoluminescence of New Europium Complex Eu(DBM)3(DPPZ) with Dipyridophenazine Ligand. Chin. Chem. Lett. 2003, 14, 43–46. [Google Scholar]
  38. Sheldrick, G.M.S. SHELXL-97—A Program for Crystal Structure Refinement; University of Goettingen: Goettingen, Germany, 1997. [Google Scholar]
  39. Ouellette, M.; Borst, P. Drug-resistance and P-glycoprotein gene amplification in the protozoan parasite Leishmania. Res. Microbiol. 1991, 142, 737–746. [Google Scholar] [CrossRef]
  40. Monte-Neto, R.L.; Coelho, A.C.; Raymond, F.; Légaré, D.; Corbeil, J.; Melo, M.N.; Frézard, F.; Ouellette, M. Gene Expression Profiling and Molecular Characterization of Antimony Resistance in Leishmania amazonensis. PLoS Negl. Trop. Dis. 2011, 5, e1167. [Google Scholar]
  41. Ma, G.Y.; Khan, S.I.; Jacob, M.R.; Tekwani, B.L.; Li, Z.Q.; Pasco, D.S.; Walker, L.A.; Khan, L.A. Antimicrobial and antileishmanial activities of hypocrellins A and B. Antimicrob. Agents Chemother. 2004, 48, 4450–4452. [Google Scholar]
  42. Habtemariam, S. In vitro antileishmanial effects of antibacterial diterpenes from two Ethiopian Premma species: P. chimperi and P. oligotricha. BMC Pharmacol. 2003, 3, 1–6. [Google Scholar] [CrossRef]
  43. Ouellette, M.; Fase-Fowler, F.; Borst, P. The amplified H circle of methotrexate-resistant leishmania tarentolae contains a novel P-glycoprotein gene. EMBO J. 1990, 9, 1027–1033. [Google Scholar]
  44. Mosmann, T. Rapid colorimetric assay for cellular growth and survival: Application to proliferation and cytotoxicity assays. J. Immunol. Methods 1983, 65, 55–63. [Google Scholar] [CrossRef]
  45. Wilson, J.J.; Lippard, S.J. In Vitro Anticancer Activity of cis-Diammineplatinum(II) Complexes with β-Diketonate Leaving Group Ligands. J. Med. Chem. 2012, 55, 5326–5336. [Google Scholar]
  • Sample Availability: Samples of the antimony(III) and bismuth(III) complexes of dppz are available from the authors.

Share and Cite

MDPI and ACS Style

Lizarazo-Jaimes, E.H.; Monte-Neto, R.L.; Reis, P.G.; Fernandes, N.G.; Speziali, N.L.; Melo, M.N.; Frézard, F.; Demicheli, C. Improved Antileishmanial Activity of Dppz through Complexation with Antimony(III) and Bismuth(III): Investigation of the Role of the Metal. Molecules 2012, 17, 12622-12635. https://doi.org/10.3390/molecules171112622

AMA Style

Lizarazo-Jaimes EH, Monte-Neto RL, Reis PG, Fernandes NG, Speziali NL, Melo MN, Frézard F, Demicheli C. Improved Antileishmanial Activity of Dppz through Complexation with Antimony(III) and Bismuth(III): Investigation of the Role of the Metal. Molecules. 2012; 17(11):12622-12635. https://doi.org/10.3390/molecules171112622

Chicago/Turabian Style

Lizarazo-Jaimes, Edgar H., Rubens L. Monte-Neto, Priscila G. Reis, Nelson G. Fernandes, Nivaldo L. Speziali, Maria N. Melo, Frédéric Frézard, and Cynthia Demicheli. 2012. "Improved Antileishmanial Activity of Dppz through Complexation with Antimony(III) and Bismuth(III): Investigation of the Role of the Metal" Molecules 17, no. 11: 12622-12635. https://doi.org/10.3390/molecules171112622

Article Metrics

Back to TopTop