Next Article in Journal
Screen Printable Sol-Gel Materials for High-Throughput Borosilicate Glass Film Production
Previous Article in Journal
Evaluating the Biodegradation of Veterinary Antibiotics Using Kinetics Model and Response Surface Methodology
Previous Article in Special Issue
Aza-Oxa-Triazole Based Macrocycles with Tunable Properties: Design, Synthesis, and Bioactivity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis, Structure and Antileishmanial Evaluation of Endoperoxide–Pyrazole Hybrids

by
Patrícia S. M. Amado
1,2,†,
Inês C. C. Costa
1,2,†,
José A. Paixão
3,
Ricardo F. Mendes
4,
Sofia Cortes
5 and
Maria L. S. Cristiano
1,2,*
1
Center of Marine Sciences, CCMAR, Gambelas Campus, University of Algarve, UAlg, 8005-139 Faro, Portugal
2
Department of Chemistry and Pharmacy, Faculty of Sciences and Technology, FCT, Gambelas Campus, University of Algarve, UAlg, 8005-139 Faro, Portugal
3
CFisUC, Department of Physics, University of Coimbra, 3004-516 Coimbra, Portugal
4
CICECO-Aveiro Institute of Materials, University of Aveiro, 3810-193 Aveiro, Portugal
5
Global Health and Tropical Medicine-GHTM, Instituto de Higiene e Medicina Tropical-IHMT, Universidade Nova de Lisboa-NOVA, 1349-008 Lisboa, Portugal
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2022, 27(17), 5401; https://doi.org/10.3390/molecules27175401
Submission received: 11 July 2022 / Revised: 19 August 2022 / Accepted: 22 August 2022 / Published: 24 August 2022

Abstract

:
Leishmaniases are among the most impacting neglected tropical diseases. In attempts to repurpose antimalarial drugs or candidates, it was found that selected 1,2,4-trioxanes, 1,2,4,5-tetraoxanes, and pyrazole-containing chemotypes demonstrated activity against Leishmania parasites. This study reports the synthesis and structure of trioxolane–pyrazole (OZ1, OZ2) and tetraoxane–pyrazole (T1, T2) hybrids obtained from the reaction of 3(5)-aminopyrazole with endoperoxide-containing building blocks. Interestingly, only the endocyclic amine of 3(5)-aminopyrazole was found to act as nucleophile for amide coupling. However, the fate of the reaction was influenced by prototropic tautomerism of the pyrazole heterocycle, yielding 3- and 5-aminopyrazole containing hybrids which were characterized by different techniques, including X-ray crystallography. The compounds were evaluated for in vitro antileishmanial activity against promastigotes of L. tropica and L. infantum, and for cytotoxicity against THP-1 cells. Selected compounds were also evaluated against intramacrophage amastigote forms of L. infantum. Trioxolane–pyrazole hybrids OZ1 and OZ2 exhibited some activity against Leishmania promastigotes, while tetraoxane–pyrazole hybrids proved inactive, most likely due to solubility issues. Eight salt forms, specifically tosylate, mesylate, and hydrochloride salts, were then prepared to improve the solubility of the corresponding peroxide hybrids and were uniformly tested. Biological evaluations in promastigotes showed that the compound OZ1•HCl was the most active against both strains of Leishmania. Such finding was corroborated by the results obtained in assessments of the L. infantum amastigote susceptibility. It is noteworthy that the salt forms of the endoperoxide–pyrazole hybrids displayed a broader spectrum of action, showing activity in both strains of Leishmania. Our preliminary biological findings encourage further optimization of peroxide–pyrazole hybrids to identify a promising antileishmanial lead.

Graphical Abstract

1. Introduction

Leishmaniases are a set of vector-borne infectious syndromes classified as neglected tropical diseases. Infections are caused by intracellular protozoans belonging to the genus Leishmania, from which more than 20 infective species are known worldwide [1,2,3]. Epidemiological data estimate 1.5 to 2 million new cases annually, with visceral leishmaniasis (VL) accounting for 500,000 cases and cutaneous leishmaniasis (CL) for 1 to 1.5 million cases [4]. Antileishmanial chemotherapy relies on a restricted arsenal of drugs comprising antimony-salts, paromomycin, miltefosine, and amphotericin B, all of which have serious shortcomings, such as unwanted side effects, unsatisfactory administration routes, and high cost and are also facing decreasing efficacy due to the emergence of resistances. These drawbacks prompted the search for better options to address leishmaniasis treatment [1,5].
Recently, it was demonstrated that selected synthetic 1,2,4,5-tetraoxanes and 1,2,4-trioxolanes exhibit antiparasitic activity against Leishmania spp. (Figure 1A) [1,5,6,7,8]. These endoperoxide classes have been extensively explored as an alternative to the natural compound artemisinin in the fight against malaria [9,10,11,12,13]. As for endoperoxides, the limelight of pyrazoles is prominent in the medicinal scope, extending their anti-inflammatory [14], antineoplastic [15], antiviral [16], and antileishmanial activity (Figure 1A) [17]. The pyrazole moiety exhibits prototropic tautomerism, leading to distinctive pyrazole-based structures with diverse reactivities [18]; so this class remains under intense scrutiny [19,20].
Bioactive endoperoxides can be coupled with other biologically active moieties, preferentially those acting through a different mechanism of action, to generate hybrids that integrate multiple pharmacophores in a single chemical entity, thus enabling the development of endoperoxide-based multi-target drugs [21]. This approach aims to increase effectiveness through a synergistic effect and to delay resistance selection. The linker between the pharmacophoric moieties can be tailored to adjust the drug’s pharmacokinetic performance and bioavailability [22,23]. Given the potential of endoperoxides and pyrazoles as tools against parasitic diseases, namely malaria and leishmaniases, we propose to explore endoperoxide–pyrazole hybrids. In this work, we disclose the synthesis and structure of 1,2,4,5-tetraoxane–pyrazole (T1, T2) and 1,2,4-trioxolane–pyrazole (OZ1, OZ2) hybrids (Figure 1B). The structures of the tetraoxane-based hybrids (5-amino-1H-pyrazol-1-yl)(dispiro[cyclohexane-1,3′-[1,2,4,5]tetraoxane-6′,2″-tricyclo[3.3.1.13,7]decan]-4-yl)methanone (T1) and (3-amino-1H-pyrazol-1-yl)(dispiro[cyclohexane-1,3′-[1,2,4,5]tetraoxane-6′,2″-tricyclo[3.3.1.13,7]decan]-4-yl)methanone (T2) were elucidated by X-ray crystallography. We conducted studies to evaluate the activity of the hybrids against Leishmania infantum, responsible for both CL and VL, and Leishmania tropica, responsible for CL [24,25], to explore novel antiparasitic candidates. By synthesizing and fully characterizing these novel compounds, we intended to explore their activity against protozoan parasites using Leishmania spp. as first models.

2. Results and Discussion

2.1. Synthesis and Structure Analysis

The novel 1,2,4-trioxolane–pyrazoles OZ1/OZ2 and 1,2,4,5-tetraoxane–pyrazoles T1/T2 were synthesized by adapting procedures described in the literature, and their complete structure elucidation was achieved by 1H, 13C{1H}, 2D NMR, X-ray crystallography, and HRMS. The synthetic strategies adopted for both classes are depicted in Scheme 1 and Scheme 2. The detailed procedures for their preparation and characterization are described in the Section 3 and Supporting Information (SI). The main difference between these syntheses resides in the method used to produce the endoperoxide pharmacophore. Formation of the 1,2,4-trioxolane ring relies on a Griesbaum coozonolysis, using adamantan-2-one O-methyloxime 1o, 4-oxocyclohexanecarboxylate, and ozone to yield the expected 1,2,4-trioxolane intermediate 2o. The adamantan-2-one O-methyl-oxime 1o employed in this procedure was previously synthesized from the reaction of adamantan-2-one with O-methylhydroxylamine in pyridine and methanol (Scheme 1). The method described by Amado et al. [26] was followed for the preparation of the precursor 1,2,4,5-tetraoxane 2t, starting with the peroxyacetalization of ethyl 4-oxocyclohexanecarboxylate with 50% (w/w) aqueous hydrogen peroxide in acetonitrile in the presence of the silica sulfuric acid catalyst (SSA) to afford the crude dihydroperoxide (DHP, 1t). Cyclocondensation of crude DHP 1t with adamantan-2-one catalysed by SSA produced the desired 1,2,4,5-tetraoxane 2t (Scheme 2). Hydrolysis of trioxolane-ester 2o or tetraoxane-ester 2t with lithium hydroxide in THF/H2O afforded the corresponding carboxylic acids, trioxolane 3o or tetraoxane 3t (Scheme 1 and Scheme 2). Endoperoxide carboxylic acids (3o or 3t) were then reacted with 3-aminopyrazole under an inert atmosphere in the presence of 1-ethyl-3-(3-(dimethylamino)propyl)carbodiimide hydrochloride (EDC·HCl) and hydroxybenzotriazole (HOBt), as coupling agents, and triethylamine, to yield a mixture of 1,2,4-trioxolane–pyrazole or 1,2,4,5-tetraoxane–pyrazole positional isomers (OZ1 and OZ2 or T1 and T2, Scheme 1 and Scheme 2).
The structures of both trioxolane and tetraoxane isomers were primarily elucidated by 1D and 2D Nuclear Magnetic Resonance (NMR), through 1H−1H correlation spectroscopy (COSY), heteronuclear single-quantum coherence (HSQC), and heteronuclear multiple-bond correlation (HMBC) NMR studies [Table 1 and Table 2 and Figures S11–S37 (SI)].
Due to conformational inversion, broad 13C signals were observed for the dispiro rings in 1,2,4,5-tetraoxanes in contrast to the sharper signals observed for the relatively inflexible 1,2,4-trioxolanes. Generally, in the tetraoxane dispiro system, the 13C{1H} peaks expand due to conformational flipping between the dispiro-cyclohexyl rings, resulting in the broadening of the peaks and a reduction in the number of carbons associated with those respective rings’ carbons. These observations are more frequent in six-membered ring endoperoxides, such as 1,2,4,5-tetraoxanes or 1,2,4-trioxanes [27]. Compared to the tetraoxane analogues, dispiro 1,2,4-trioxolanes exhibit much improved signal resolution and sharpness and a higher signal number in the non-aromatic region. As expected, a strong correlation coefficient was obtained in the 1H−1H COSY and HMBC spectra for endoperoxide containing dispiro rings. Furthermore, the combination of the 1H−1H COSY and HMBC spectra allowed for the complete identification of the pyrazole moiety’s carbons, enabling a more precise identification of the multiple -NH protons. Regarding the pyrazole ring, in the 1H−1H COSY spectrum, 3J correlations between H4 with H5 were observed, together with two long-range couplings with one of the -NH proton (δ = 5.54 ppm), including a four-bond coupling (4J) with H4 and a five-bond coupling (5J) with H5 in both isomers OZ1/T1 (Figures S14 and S25, SI). HMBC 3JC-NH correlations between each -NH (δ = 5.19 and 5.54 ppm) with C4 demonstrated that both isomers OZ1 and T1 have two signals in C4, implying the presence of two non-equivalent -NH protons. No HMBC 3JC-NH correlation was observed between the carbonyl carbon and the -NH from the pyrazole ring, suggesting that the exocyclic amino group did not participate in the amide generation in isomers OZ1/T1 (Figures S16 and S27, SI). These correlations indicate that a prototropic annular tautomerism [18,28] occurred during the coupling reaction, where isomers OZ1/T1 are the products from reaction of the peroxide carboxylic acids with the 5-aminopyrazole tautomer. Protons -NH are described as non-equivalent, implying a strong intramolecular hydrogen bond involving one of the exocyclic -NH protons.
For isomers OZ2 and T2, a few 1H−1H COSY correlations on the pyrazole ring are shown, with a 3J correlation between H4 and H5 (Figures S19 and S31, SI). A five-bond coupling (5J) correlation is observed between H5 and -NH2 in isomer T2 (Figure S31, SI). Regarding HMBC spectrum, a single signal 3JC-NH in C3-NH2 to C4 is observed, which demonstrates that both -NH protons are equivalent. A weak 2JC-H correlation in C3-NH2 to C3 was also observed in isomer T2 (Figure S33, SI). Notably, the data demonstrate the presence of the conjugate in its 3-aminopyrazole-linked tautomeric form, which is identifiable in the crystal phase. Theoretically, the frame of 3(5)-aminopyrazole is also susceptible to side-chain tautomerism, as it has a proton-exchangeable amine side chain [18,28], which can be observed in this situation.

2.2. X-ray Crystal Analysis

A successful preparation of the single crystals of both tetraoxane–pyrazole hybrids enabled us to further confirm the different isomers through X-ray crystallography.
The crystal of T1 used for the data collection turned out to be a non-merohedral twin corresponding to a 180° rotation around the (100) reciprocal lattice axis; the twin-law (1 0 0.860; 0 −1 0; 0 0 −1) was found using the TwinRotMax routine implemented in PLATON [29]. The proportion of the two components of the twin was found to be 59%:41% via refinement of the BASF parameter. In compound T2, examination of the residual electron density disclosed that the tetraoxane group is slightly disordered (3%) over two alternate chair conformations, which was considered in the final refinement. The same amount of disorder was observed in the analysis of the dataset collected at room temperature. Oak Ridge Thermal Ellipsoid Plots (ORTEP) of the isomer molecules of T1 and T2 are depicted in Figure 2 and Figure 3, respectively.
Bond lengths and valency angles are unexceptional, the average values in the tetraoxane ring being C–O (I: 1.429 Å; II: 1.430 Å), O–O, (I: 1.474 Å; II: 1.476 Å) and C–O–O (I: 107.8; II: 107.1°), respectively. In T2, the cyclohexane and tetraoxane rings adopt both an almost ideal chair form, with Cremer and Pople [30] puckering parameters Q: ~0.6 Å and pseudo-rotation angle θ < 2°. In the case of T1, the puckering amplitudes of these rings are similar to those of T2, and the conformation of the tetraoxane ring is also close to the ideal chair form, with θ = 1.7(6)°. However, the cyclohexane ring in T1 features a significant distortion from the ideal chair conformation, with θ = 11.3(10)° For both isomers, the substituent at C14 is at an equatorial position to the cyclohexane ring. The conformation of the adamantane substituent is close to the ideal boat–boat form of the forming eight-membered rings, with approximate Cs symmetry, in the two crystals. The pyrazole ring is planar within experimental error in both isomers. The C=O group is not strictly coplanar with the pyrazole ring, with a more pronounced rotation around the single N1–C17 bond in T2, as shown by the torsion angles O5–C17–N1–N2 (I: −175.9(8)°; II 167.90(12)°).
In the crystal of T1, one of the two H atoms of the amino group forms a short intramolecular hydrogen bond with the bare O5 atom as acceptor (Table 3). The other H atom establishes a longer intermolecular hydrogen bond with the bare N atom of the pyrazole ring of a neighbor molecule, this hydrogen bond linking the molecules in infinite chains propagating along the crystallographic c-axis (Figure 4). In the crystal structure of T2, both H atoms of the amino group are involved in intermolecular hydrogen bonds; one of them is an N–HN bond linking molecules across inversion centers in the unit cell; the other one is directed towards atom O5, as in the crystal of T1. In T2, the two hydrogen bonds form an infinite 2D network of molecules lying in the (100) plane (Figure 5). In both crystals, a set of C–HO and C–HN short contacts can be spotted, as well as C–HCg interactions with the π electron cloud of the pyrazole ring, that contribute to the stabilization of the crystal structures.

2.3. In Vitro Susceptibility of Leishmania spp. to Endoperoxide–Pyrazole Hybrids

We undertook a preliminary in vitro evaluation of the antileishmanial activity of the endoperoxide–pyrazole hybrids OZ1, OZ2, T1, T2, and 3-aminopyrazole (PYR) against promastigote forms of the viscerotropic species L. infantum and the dermotropic species L. tropica. Axenic Leishmania promastigotes were used as a model for the initial compound screening, since they are a more straightforward and economical model due to their ease of culture development, requiring small amounts of the tested molecule [31]. The half-inhibitory concentration (IC50) of each compound was determined upon 48 h incubation with each parasite strain (Table 4). As control, the standard antileishmanial drug amphotericin B (AmB) was used. Compounds cytotoxicity was assessed (CC50) on the monocytic THP-1 cell line upon exposure to each compound, and the ClogP and ClogS values were calculated (Table 4). Materials and methodologies applied in the biological evaluation of the compounds are described in the Section 3.
From the endoperoxide–pyrazole hybrids tested against L. tropica promastigotes, trioxolanes OZ1 and OZ2 displayed the lowest IC50 values (219 ± 25 µM and 161 ± 19 µM, respectively). Regarding the assessments performed on L. infantum promastigotes, this strain only revealed susceptibility towards OZ2, which exhibited an IC50 of 219 ± 44 µM, yet slightly higher than that obtained for the parent LC67 (2o) (123 ± 1 µM, Table 4). Since both ozonides have a similar value of ClogS, the lack of activity of OZ1 in L. infantum could be related to steric effects. Disparity in response to standard CL treatments or experimental drugs has become a significant burden to disease management and variations in treatment outcomes are ascribed to a number of factors, including specific characteristics and susceptibility among Leishmania species [32]; L. tropica and L. infantum are phylogenetically and genetically distinct, which justifies the appearance of different clinical forms as well as variable pharmacological susceptibility among species [33]. The calculated partition coefficient (ClogP) values proved very similar among the endoperoxides studied (1.79–2.61), suggesting that other physical properties could explain the differences in activity.
Overall, both ozonides demonstrated some antileishmanial activity, though lower than disclosed from previous in vitro studies where selected endoperoxides revealed antileishmanial potency at low micromolar (<10 µM) concentrations [5,34]. In the present study, the evaluation of compounds T1 and T2 was limited by their low solubility in M199 and RPMI medium. Calculated intrinsic aqueous solubility (ClogS) values show similar values between trioxolanes (1.20–1.38) and tetraoxanes (1.18–1.21). However, the poor solubility of dispiro 4′’-substituted tetraoxanes may be associated to their structural symmetry [35]. A decrease in molecular symmetry has been correlated to better physicochemical properties, such as absorption and solubility, resulting in enhanced systemic exposure to a drug. So it is anticipated that 3′’-substituted analogues could overcome this solubility issue [35,36]. Therefore, future optimization to 3”-substituted analogues should be considered. Likewise, conversion of the compounds into water-soluble salts, as well as encapsulation of the endoperoxide structures in delivery systems such as liposomes, could represent alternative strategies to solve the solubility issue [37,38].
Cytotoxicity tests revealed toxic effects of the hybrids OZ1 and OZ2 on THP-1 cells, with low selectivity (SI values of <5 µM). We observed that the parent 3-aminopyrazole (PYR) is devoid of antileishmanial activity but exhibits similar toxicity to OZ2, indicating that the toxicity in OZ1 and OZ2 is likely due to the aminopyrazole moiety. Moreover, the efficacy of PYR in inhibiting THP-1 (202 ± 10 µM), a monocytic leukemia human-derived cell line, also indicates that it could be a promising framework in the design of compounds with antineoplastic potential.

2.4. Effects of Endoperoxide–Pyrazole Hybrids on L. tropica Promastigotes Morphology

To evaluate the effect promoted by endoperoxide–pyrazole hybrids on Leishmania morphology, which can impair its infection capacity, we performed a morphometric analysis with optical microscopy (1000× amplification) in L. tropica promastigotes treated for 48 h with 300 µM of OZ1 and 200 µM of OZ2 as these concentrations were close to the IC50 values obtained for this species (219 ± 25 and 161 ± 19 µM, respectively).
Figure 6 illustrates the measurements of L. tropica body (A) and flagellum (B) of parasites treated with both compounds and the control drug amphotericin B in comparison with non-treated parasites. Results show significant differences p < 0.001 for both compounds. Despite exhibiting a higher IC50 than OZ2, compound OZ1 induced a greater reduction in body size (5.265 ± 1.733 µm) than OZ2 (9.646 ± 2.744 µm), even slightly more pronounced than amphotericin B (6.463 ± 1.874 µm). On the other hand, both molecules decreased the flagellum size, with effects within the same range for OZ2 (12.404 ± 3.745 µm) and OZ1 (13.561 ± 3.535 µm). In general, changes in flagellum size are less evident when compared to body size loss; nevertheless, they may limit mobility and its capacity in infecting cells.
Representative images of cellular alterations are depicted in Figure 7C–F. Non-treated cells and amphotericin B-treated cells were used as negative and positive controls, respectively (Figure 7A,B).
After 48 h of incubation, non-treated promastigotes (Figure 7A) exhibited a fusiform-shaped body, well-defined intracellular organelles, and an elongated flagellum. Contrarywise, in cultures treated with OZ1 (Figure 7C,D) and OZ2 (Figure 7E,F), most promastigotes showed considerable damage, namely significant body size reduction with an abnormally rounded shape, cytoplasmic disintegration, apparent disruption of internal organelles, and cell shrinkage, which are the typical alterations observed upon treatment with established antileishmanial compounds [39]. Another parallel impact of these compounds is the flagellum length reduction as discussed previously, which may cause impairments in motility. As expected, all these constraints were also observed in promastigotes treated with amphotericin B (Figure 7B).
Even though the IC50 values indicated that the ozonides have some activity, microscopic analysis revealed that L. tropica promastigotes experienced cell changes analogous to those caused by amphotericin B. These findings could support further investigations on the endoperoxide–pyrazole hybrids in the frame of leishmaniases chemotherapy, underlining the need for structure–activity studies and further optimisation.

2.5. Evaluation of the Hybrids’ Stability upon Conversion to Their Salt Forms

The antileishmanial activity of the hybrids on promastigotes of L. tropica revealed the need for structural modifications to improve the antileishmanial effect of these compounds. Because solubility may play a significant role in the antileishmanial activity of this series of molecules, we converted the free amines attached to the pyrazole moiety into their salt forms. Several acids, namely p-toluenesulfonic, methanesulfonic, and hydrochloric acid, can generate the free amines into salts with variable solubility levels. As such, the endoperoxide–pyrazole hybrids OZ1, OZ2, T1 and T2 were converted into their p-tosylate (OZ1•TsOH, OZ2•TsOH, and T1•TsOH), mesylate (OZ1•MsOH), and hydrochloride (OZ1•HCl, OZ2•HCl, T1•HCl, and T2•HCl) salts, with low to good yields (Scheme 3). Interestingly, this study has shown that the endoperoxide–pyrazole hybrids are stable even in strongly acidic conditions without decomposition of the endoperoxide ring, allowing further structure optimization.
Tests of antileishmanial activity were conducted on promastigote forms of L. tropica and L. infantum, then also on intracellular amastigote forms of L. infantum to ascertain the effect of the endoperoxide–pyrazole hybrids in their salt forms. The results are depicted in Table 5. The different salts displayed different antileishmanial efficacy depending on the conjugate salt, with IC50 values ranging from 135– > 400 µM against L. tropica, and from 164– > 400 µM against L. infantum. Among the salt forms of endoperoxide–pyrazole hybrids tested against L. tropica promastigotes the trioxolane OZ1HCl displayed an IC50 value of 135 ± 36 µM, indicating that different salts of the same isomer exhibit divergent antileishmanial activities, with the OZ1•HCl salt proving to be the most active salt against L. tropica. Unlike for the results obtained for OZ1, OZ2, T1 and T2, the assessments performed on L. infantum promastigotes revealed susceptibility of this strain towards most of the different salts except for the T1 analog with TsOH and HCl. As observed for L. tropica, the trioxolane OZ1HCl also displayed the lowest IC50 value (164 ± 20 µM) against L. infantum. Cytotoxicity assessments revealed toxic effects of all the salt derivatives on THP-1 cells, with CC50 values compressed in a concentration range of 331–529 µM. The improvements in solubility exhibited by the different salts enhanced the antileishmanial activity, although further structure–activity optimizations are still needed. However, it is worth noting that the salt forms of the endoperoxide–pyrazole hybrids display a broader spectrum of action, showing some effect on both strains of Leishmania.
Compounds that exhibited an IC50 < 200 µM were also tested for intracellular Leishmania amastigotes susceptibility. The L. infantum strain was used as the model for this assay. According to the results of intracellular amastigote susceptibility (Table 4 and Table 5), OZ1•HCl and IC0 (2t) demonstrated superior activity compared to the other compounds, with lower IC50 values against intracellular amastigotes (87 ± 39 µM and 45 ± 28 µM) than in promastigotes (164 ± 20 µM and 128 ± 27 µM). In contrast, OZ1•TsOH, OZ1•MsOH and LC67 evidenced a diminished effect on amastigotes compared to that shown on promastigotes (Figure 8). Possible reasons could be linked to bioaccumulation and mechanism of action for this class of compounds, including passage through the host cell’s cellular membrane or the formation of toxic metabolites toward Leishmania after metabolization of these compounds by host cell enzymes. As reported, the parasitophorous vacuole may serve as a barrier to direct interaction with the parasite, which was reported to eventually lead to a decrease in activity on moving from promastigote to amastigote stages [40]. Thus, it is noteworthy that OZ1•HCl and IC0 (2t) demonstrated superior activity compared to the other compounds, with slightly lower IC50 values against intracellular amastigotes than in promastigotes. Overall, the compounds tested against the amastigote forms displayed poor selectivity since SI values are smaller than 10.

3. Materials and Methods

3.1. Chemicals

All reagents for synthesis were purchased from commercial sources and used without further purification. Analytical thin layer chromatography (TLC) was carried out using Merck (Darmstadt, Germany) TLC Silica gel 60 F254 aluminum sheets and visualized under UV or by appropriate stain. P-Anisaldehyde and potassium permanganate were the most used. Column chromatography was carried out using Sigma Aldrich (Darmstadt, Germany) technical grade silica gel (pore size 60 Å, 230–400 mesh particle size, 40–63 μm particle size).

3.2. Equipment

1H and 13C Nuclear Magnetic Resonance (NMR) spectra were recorded using a 400 MHz Bruker (Billerica, MA, USA) instrument or using a 500 MHz JEOL system (Peabody, MA, USA) equipped with a Royal HFX probe in the deuterated solvents described in each experimental procedure. The chemical shifts (δ) are described in parts per million (ppm) downfield from an internal standard of tetramethylsilane (TMS). Melting points (°C) were obtained on an SMP30 melting point apparatus and are uncorrected. High-Resolution Mass Spectrometry (HRMS) was recorded using the analytical service within the Centre of Marine Sciences (CCMAR, Algarve, Portugal) and was conducted on a Thermo Scientific High Resolution Mass Spectrometer (HRMS) (Waltham, MA, USA), model Orbitrap Elite, capable of MSn, n up to 10. X-ray diffraction data were collected on Bruker Kappa Apex II and D8 Quest diffractometers (Billerica, MA, USA) using graphite monochromated MoKα (λ = 0.71073 Å) radiation (see Section 3.4 for more details).
Safety. Organic peroxides are potentially hazardous compounds (inflammable and explosive) and must be handled carefully: (1) a safety shield should be used for all reactions involving H2O2; (2) direct exposure to strong heat or light, mechanical shock, oxidizable organic materials, or transition-metal ions should be avoided.

3.3. Synthesis

General Procedure 1: Removal of the ester group. Followed procedure by Jiricek et al. with slight modifications [41]. To a solution of peroxide-ester (1 mmol) in THF:H2O (1:1, 5 mL), LiOH (3 mmol) was added, and the reaction mixture was stirred until completion, at room temperature. THF was removed in vacuo, diluted with water, and acidified with 1M HCl. The aqueous layer was extracted with EtOAc (3 × 30 mL), washed with brine (2 × 30 mL), and dried with anhydrous MgSO4. The organic phase was then filtered and concentrated under reduced pressure. The solid obtained was then crushed and washed with cold n-hexane, giving the desired product.
General Procedure 2: Synthesis of the peroxide–pyrazole hybrids. Followed the procedure by Stec et al. [42] with slight modifications. Peroxide–carboxylic acid (1 mmol) was dissolved in anhydrous dichloromethane (DCM, 5 mL) at room temperature. To the solution, was added hydroxybenzotriazole (HOBt, 1.5 mmol) and 1-ethyl-3-(3-(dimethylamino)propyl)-carbodiimide hydro-chloride (EDC·HCl, 1.5 mmol) under a nitrogen atmosphere. After stirring for 1 h, 3-aminopyrazole (1.1 mmol) and triethylamine (2 mmol) were added, and the reaction mixture was stirred at room temperature until consumption of the starting material (usually overnight). To the resulting solution was added water, and the mixture was extracted with DCM (3 × 15 mL) and EtOAc (3 × 15 mL). The organic layers were separated, washed with brine, dried over anhydrous MgSO4, filtered, and concentrated under reduced pressure. The residue was purified by flash chromatography using a 0–40% EtOAc−hexane gradient (unless specified differently) to obtain the peroxide–pyrazole hybrids.
General Procedure 3: Synthesis of the peroxide-pyrazole p-tosylate salts. Procedure adapted from Vennerstrom et al. [43]. A solution of p-toluenesulfonic acid monohydrate (0.9 mmol) in ethanol (2.5 mL) was added to a solution of the endoperoxide–pyrazole hybrid (1 mmol) in CHCl3 (10 mL). The product formed was concentrated under reduced pressure and treated with diethyl ether (20 mL). The precipitate was then filtered, washed with diethyl ether (20 mL), and dried to yield the desired endoperoxide–pyrazole p-tosylate salt.
General Procedure 4: Synthesis of the peroxide–pyrazole mesylate salts. Procedure adapted from Vennerstrom et al. [43]. A solution of methanesulfonic acid (0.73 mmol) in diethyl ether (2.5 mL) was added to a solution of the endoperoxide–pyrazole hybrid (1 mmol) in CHCl3 (5 mL). The precipitate was then filtered, washed with diethyl ether (20 mL), and dried to yield the desired endoperoxide–pyrazole mesylate salt.
General Procedure 5: Synthesis of the peroxide–pyrazole hydrochloride salts. Procedure adapted from Vennerstrom et al. [43]. Anhydrous HCl–Et2O (2.0 M, 10 mL was added to a solution of the endoperoxide–pyrazole hybrid (1 mmol) in CHCl3 (5 mL). The precipitate was then filtered, washed with diethyl ether (20 mL), and dried to yield the desired endoperoxide–pyrazole hydrochloride salt.
O-methyl-adamantan-2-one oxime (1o). Followed procedure by Cabral et al. [1] with slight modifications. To a solution of adamantan-2-one (7.00 g, 46.6 mmol) in methanol (50 mL) were added pyridine (6.00 mL, 74.18 mmol) and methoxylamine hydrochloride (5.06 g, 60.52 mmol). The reaction mixture was stirred at room temperature for 48 h. The final mixture was concentrated and then diluted with DCM (50 mL) and water (50 mL). The organic layer was separated, and the aqueous layer was extracted with DCM (30 mL). The combined organic extracts were washed with aqueous HCl (1 M; 30 mL × 2), then with brine (30 mL). The final organic extract was dried over MgSO4, filtered, and concentrated under reduced pressure to give O-methyl-adamantan-2-one oxime (6.01 g, 72% yield) as a colourless solid. M.p. 69–70 °C. Spectral data are in accordance with those reported [1]. 1H NMR (400 MHz, CDCl3): δ 1.69–2.02 (m, 14H), 2.14 (t, J = 6.9 Hz, 4H), 2.51 (t, J = 7.0 Hz, 4H) ppm; 13C{1H} NMR (100 MHz, CDCl3): δ 25.9, 26.31, 31.09, 32.59, 34.25, 35.70, 36.18, 37.35, 106.46, 111.95, 208.90 ppm. HRMS (ES+, m/z) calcd C11H18NO (M+H)+: 180.13829; found 180.13783. Diff: 2.55 ppm.
Ethyl dispiro[cyclohexane-1,3′-[1,2,4]trioxolane-5′,2″-tricyclo[3.3.1.13,7]decane]-4-carboxylate (2o). Followed procedure by Cabral et al. [1] with slight modifications. Ozone, produced with an ozone generator Sander Labor-Ozonizator 301.7 (0.5 L/min O2, 140 V), which was passed through a solution of dichloromethane at –78 °C and flushed into a solution of 1o (2.00 g, 23.95 mmol) and ethyl 4-oxocyclohexanecarboxylate (1.89 mL, 11.97 mmol), in pentane/DCM (6:4, 100 mL) at 0 °C. After completion, the solution was flushed with nitrogen for 5 min and concentrated under reduced pressure at room temperature to give a crude material that was purified by column chromatography. (EtOAc: n-hexane, 1:99, v/v), affording a colourless oil (1.73 g, 48% yield). Spectral data are in accordance with those reported [1]. 1H NMR (500 MHz, CDCl3): δ 4.11 (dq, J = 9.4, 7.1 Hz, 2H), 2.40–2.26 (m, 1H), 2.00–1.66 (m, 22H), 1.24 (dt, J = 8.3, 7.1 Hz, 3H). 13C{1H} NMR (126 MHz, CDCl3): δ 175.04, 111.60, 108.04, 60.49, 41.52, 41.30, 37.15, 36.85, 36.44, 34.94, 34.90, 34.85, 34.83, 34.06, 33.66, 33.44, 33.29, 33.23, 31.59, 27.17, 27.02, 26.94, 26.53, 26.31, 26.13, 14.33, 14.30. HRMS (ES+, m/z) calcd C19H32NO5 (M+NH4)+: 354.22750; found 354.22702. Diff: 1.36 ppm.
Ethyl dispiro[cyclohexane-1,3′-[1,2,4,5]tetraoxane-6′,2″-tricyclo[3.3.1.13,7]decane]-4-carboxylate (2t). Procedure adapted from Amado et al. [26]. Ethyl 4-oxocyclohexanecarboxylate (2 mL, 12.55 mmol) was dissolved in acetonitrile (38 mL), and silica sulfuric acid (SSA, 3.42 g, 2 mmol) was added to the mixture. Hydrogen peroxide 50 wt.% in H2O (4.09 mL, 4 mmol) was slowly added over an ice bath. Then the mixture was left to stir at room temperature until consumption of the starting material. To this mixture was added distilled water. Then the catalyst was filtered and rinsed with DCM. The filtrate was extracted with DCM (3 × 30 mL), dried over with MgSO4, and concentrated under reduced pressure at low temperature (30–35 °C) to obtain the gem-dihydroperoxide semi-crude, which was used immediately without further purification. The gem-dihydroperoxide semi-crude (1t) was dissolved in anhydrous DCM (75 mL) followed by addition of 2-adamantanone (5.66 g, 1.5 mmol). The mixture was cooled over an ice bath prior to addition of SSA (3.42 g, 2 mmol). The mixture was then warmed and left to stir at room temperature until consumption of the starting material. The resulting solution was then filtered, rinsed with DCM, and concentrated under reduced pressure. Purification through flash chromatography (EtOAc: n-hexane, 2.5:97.5, v/v) yielded a white solid (2.11 g, 48% yield). M.p. = 67–69 °C. Spectral data are in accordance with those reported [26]. 1H NMR (500 MHz, CDCl3): δ 4.12 (q, J = 7.1 Hz, 2H), 3.02 (d, J = 118.6 Hz, 2H), 2.41–2.34 (m, 1H), 2.08–1.60 (m, 19H), 1.50 (s, 1H), 1.23 (t, J = 7.1 Hz, 3H). 13C{1H} NMR (126 MHz, CDCl3): δ 174.75, 110.61, 107.28, 60.53, 41.77, 39.35, 37.01, 34.35, 33.21, 30.18, 30.16, 28.29, 27.11, 24.82, 23.94, 14.31. HRMS (ESI+, m/z) calcd C19H28O6Na (M+Na)+: 375.17781; found 375.17725. Diff: 1.49 ppm.
Dispiro[cyclohexane-1,3′-[1,2,4]trioxolane-5′,2″-tricyclo[3.3.1.13,7]decane]-4-carboxylic acid (3o). This compound was synthesized in accordance with general procedure 1, using 2o. White solid (2.01 g, 88% yield). M.p. = 148–150 °C. Spectral data are in accordance with those reported [44]. 1H NMR (500 MHz, CDCl3): δ 2.43–2.33 (m, 1H), 2.00–1.67 (m, 22H). 13C{1H} NMR (126 MHz, CDCl3): δ 181.06, 180.94, 111.93, 111.68, 107.93, 107.88, 41.06, 40.85, 39.36, 36.84, 36.43, 34.95, 34.89, 34.85, 34.82, 33.30, 33.06, 27.17, 26.93, 26.52, 26.05, 25.88. HRMS (ESI, m/z) calcd C17H23O5 (M−H): 307.15510; found 307.15488. Diff: 0.72 ppm.
Dispiro[cyclohexane-1,3′-[1,2,4,5]tetraoxane-6′,2″-tricyclo[3.3.1.13,7]decane]-4-carboxylic acid (3t). This compound was synthesized in accordance with general procedure 1, using 2t. White solid (1.96 g, 86% yield). M.p. = 179–182 °C. Spectral data are in accordance with those reported [1]. 1H NMR (500 MHz, CDCl3): δ 3.14 (s, 1H), 2.91 (s, 1H), 2.49–2.42 (m, 1H), 2.06–1.72 (m, 17H), 1.62–1.44 (m, 3H). 13C{1H} NMR (126 MHz, CDCl3): δ 180.57, 110.66, 107.12, 41.32, 39.36, 37.01, 34.33, 33.30, 33.21, 32.31, 30.26, 28.21, 27.12, 24.58, 23.68. HRMS (ESI, m/z) calcd C17H23O6 (M−H): 323.15001; found 323.14975. Diff: 0.80 ppm.
(5-Amino-1H-pyrazol-1-yl)(dispiro[cyclohexane-1,3′-[1,2,4]trioxolane-5′,2″-tricyclo-[3.3.1.13,7]-decan]-4-yl)methanone (OZ1) and (3-Amino-1H-pyrazol-1-yl)(dispiro[cyclohexane-1,3′-[1,2,4]trioxolane-5′,2″-tricyclo[3.3.1.13,7]decan]-4-yl)methanone (OZ2). The compounds were prepared in accordance with general procedure 2, through amide coupling of 3o with 3(5)-aminopyrazole. Purification by flash chromatography (EtOAc: n-hexane, 20:80, v/v) yielded the two isomers as white solids (OZ1: 0.81 g, 28% yield; OZ2: 1.22 g, 42%).
OZ1: M.p. = 153–155 °C. TLC Rf: 0.41 (EtOAc: n-hexane, 20:80, v/v). 1H NMR (500 MHz, CDCl3): δ 7.37 (d, J = 1.8 Hz, 1H), 5.54 (s, 2H), 5.37 (dd, J = 4.6, 1.8 Hz, 1H), 3.59 (dtq, J = 11.7, 9.3, 3.6 Hz, 1H), 1.87 (m, 22H). 13C{1H} NMR (126 MHz, CDCl3): δ 178.50, 150.35, 144.01, 111.64, 108.00, 88.93, 40.86, 36.87, 36.47, 34.87, 34.83, 33.46, 26.95, 26.55, 26.34. HRMS (ES+, m/z) calcd C20H28N3O4 (M+H)+: 374.20743; found 374.20706. Diff: 0.99 ppm.
OZ2: M.p. = 64–66 °C. TLC Rf: 0.23 (EtOAc: n-hexane, 20:80, v/v). 1H NMR (500 MHz, CDCl3): δ 7.99 (d, J = 3.0 Hz, 1H), 5.88 (d, J = 3.0 Hz, 1H), 3.99 (s, 2H), 3.45–3.36 (m, 1H), 2.06–1.71 (m, 22H). 13C{1H} NMR (126 MHz, CDCl3): δ 173.12, 157.34, 130.13, 130.09, 111.98, 111.58, 108.06, 108.03, 101.03, 100.97, 39.64, 36.87, 36.47, 36.45, 34.95, 34.89, 34.87, 34.83, 33.51, 33.43, 26.96, 26.55, 26.45, 26.36. HRMS (ES+, m/z) calcd C20H28N3O4 (M+H)+: 374.20743; found 374.20752. Diff: -0.24 ppm.
(5-Amino-1H-pyrazol-1-yl)(dispiro[cyclohexane-1,3′-[1,2,4,5]tetraoxane-6′,2″-tricyclo-[3.3.1.13,7]-decan]-4-yl)methanone (T1) and (3-Amino-1H-pyrazol-1-yl)(dispiro[cyclohexane-1,3′-[1,2,4,5]tetraoxane-6′,2″-tricyclo[3.3.1.13,7]decan]-4-yl)methanone (T2). The compounds were synthesised in accordance with general procedure 2, through amide coupling of 3t with 3(5)-aminopyrazole. Purification by flash chromatography (EtOAc: n-hexane, 20:80, v/v) yielded the two isomers as white solids (T1: 0.35 g, 23% yield; T2: 0.49 g, 31%).
T1: M.p. = 187–189 °C (dec.). TLC Rf: 0.40 (EtOAc: n-hexane, 20:80, v/v). 1H NMR (500 MHz, CDCl3): δ 7.36 (d, J = 1.8 Hz, 1H), 5.54 (s, 2H), 5.37 (d, J = 1.8 Hz, 1H), 3.66 (ddt, J = 11.2, 7.5, 3.8 Hz, 1H), 3.17 (s, 2H), 2.11–1.57 (m, 22H). 13C{1H} NMR (126 MHz, CDCl3): δ 178.17, 150.34, 144.04, 110.65, 107.22, 88.95, 41.32, 37.02, 34.05, 33.23, 30.24, 28.40, 27.13, 24.85, 24.44. HRMS (ES+, m/z) calcd C20H28N3O5 (M+H)+: 390.20235; found 390.20181. Diff: 1.38 ppm.
T2: M.p. = 171–173 °C (dec.). TLC Rf: 0.22 (EtOAc: n-hexane, 20:80, v/v). 1H NMR (500 MHz, CDCl3): δ 8.03–7.97 (d, J = 2.9 Hz, 1H), 5.90–5.87 (d, J = 3.0 Hz, 1H), 4.02–3.89 (s, 2H), 3.54–3.45 (tt, J = 11.1, 3.8 Hz, 1H), 3.28–3.07 (s, 2H), 2.05–1.61 (m, 20H). 13C{1H} NMR (126 MHz, CDCl3): δ 178.17, 150.34, 144.04, 110.65, 107.22, 88.95, 41.32, 37.02, 34.05, 33.23, 30.24, 28.40, 27.13, 24.85, 24.09. HRMS (ES+, m/z) calcd C20H28N3O5 (M+H)+: 390.20235; found 390.20206. Diff: 0.74 ppm.
(5-Amino-1H-pyrazol-1-yl)(dispiro[cyclohexane-1,3′-[1,2,4]trioxolane-5′,2″-tricyclo-[3.3.1.13,7]-decan]-4-yl)methanone p-tosylate (OZ1•TsOH). The compound was prepared from OZ1, in accordance with general procedure 3. White solid (87 mg, 34% yield). M.p. = 149 °C (dec.). 1H NMR (500 MHz, DMSO-d6): δ 7.51–7.43 (d, J = 8.0 Hz, 2H), 7.44–7.37 (d, J = 1.8 Hz, 1H), 7.21–7.06 (d, J = 7.8 Hz, 2H), 5.34–5.27 (d, J = 1.8 Hz, 1H), 3.57–3.44 (ddt, J = 11.8, 7.1, 3.7 Hz, 1H), 2.32–2.25 (s, 3H), 2.05–1.40 (m, 22H) ppm. 13C{1H} NMR (126 MHz, DMSO-d6): δ 177.47, 151.10, 145.59, 144.28, 137.78, 128.15, 125.54, 110.83, 107.87, 87.33, 40.11, 40.02, 36.10, 35.78, 34.28, 32.78, 26.23, 25.97, 25.83, 20.83 ppm.
(5-Amino-1H-pyrazol-1-yl) (dispiro[cyclohexane-1,3′-[1,2,4]trioxolane-5′,2″-tricyclo-[3.3.1.13,7]-decan]-4-yl) methanone mesylate (OZ1•MsOH). The compound was prepared in accordance with general procedure 4, using OZ1. White solid (110 mg, 62% yield). M.p. = 118 °C (dec.). 1H NMR (500 MHz, DMSO-d6): δ 7.43–7.38 (d, J = 1.8 Hz, 1H), 5.34–5.28 (d, J = 1.8 Hz, 1H), 3.59–3.51 (tt, J = 11.7, 3.4 Hz, 1H), 2.41–2.38 (s, 3H), 1.95–1.53 (m, 22H) ppm. 13C{1H} NMR (126 MHz, DMSO-d6): δ 177.47, 151.10, 144.28, 110.84, 107.88, 87.34, 40.11, 40.02, 36.11, 35.78, 34.29, 32.78, 26.24, 25.98, 25.84 ppm.
(5-Amino-1H-pyrazol-1-yl) (dispiro[cyclohexane-1,3′-[1,2,4]trioxolane-5′,2″-tricyclo-[3.3.1.13,7]-decan]-4-yl) methanone hydrochloride (OZ1•HCl). The compound was prepared in accordance with general procedure 5, using OZ1. White solid (53 mg, 27% yield). 1H NMR (500 MHz, DMSO-d6): δ 7.41 (d, J = 1.8 Hz, 1H), 5.31 (dd, J = 3.2, 1.8 Hz, 1H), 3.55 (tt, J = 11.7, 3.4 Hz, 1H), 1.96–1.53 (m, 24H) ppm. 13C{1H} NMR (126 MHz, DMSO-d6): δ 177.46, 151.09, 144.26, 110.82, 107.87, 87.32, 36.09, 35.77, 34.44, 34.27, 32.77, 32.71, 26.22, 25.96, 25.82 ppm.
(3-Amino-1H-pyrazol-1-yl) (dispiro[cyclohexane-1,3′-[1,2,4]trioxolane-5′,2″-tricyclo-[3.3.1.13,7] decan]-4-yl) methanone p-tosylate (OZ2•TsOH). The compound was synthesised in accordance with general procedure 3, using OZ2. White solid (89 mg, 23% yield). M.p. = 129 °C (dec.). 1H NMR (500 MHz, DMSO-d6): δ 8.03 (d, J = 3.0 Hz, 1H), 7.48 (d, J = 8.1 Hz, 2H), 7.12 (d, J = 7.8 Hz, 2H), 5.94 (d, J = 3.0 Hz, 1H), 3.41–3.33 (m, 1H), 2.29 (s, 3H), 1.94–1.54 (m, 22H) ppm. 13C{1H} NMR (126 MHz, DMSO-d6): δ 173.16, 158.43, 145.61, 137.77, 129.61, 128.15, 125.54, 110.85, 107.85, 102.12, 38.61, 36.10, 35.77, 34.28, 32.84, 26.23, 26.09, 25.83, 20.83 ppm.
(3-Amino-1H-pyrazol-1-yl) (dispiro[cyclohexane-1,3′-[1,2,4]trioxolane-5′,2″-tricyclo-[3.3.1.13,7] decan]-4-yl) methanone hydrochloride (OZ2•HCl). The compound was synthesised in accordance with general procedure 5, using OZ2 (38 mg, 20% yield). 1H NMR (500 MHz, DMSO-d6): δ 7.89 (d, J = 2.6 Hz, 1H), 5.98 (dd, J = 3.0, 1.1 Hz, 1H), 3.47–3.22 (m, 1H), 1.89–1.66 (m, 24H) ppm. 13C{1H} NMR (126 MHz, DMSO-d6): δ 176.01, 158.40, 129.57, 110.69, 107.98, 102.15, 36.10, 35.77, 34.41, 34.30, 34.27, 32.84, 32.75, 32.60, 26.23, 26.08, 25.91, 25.83 ppm.
(5-Amino-1H-pyrazol-1-yl) (dispiro[cyclohexane-1,3′-[1,2,4,5] tetraoxane-6′,2″-tricyclo-[3.3.1.13,7]-decan]-4-yl) methanone p-tosylate (T1•TsOH). The compound was synthesised in accordance with general procedure 3, using T1. White solid (60 mg, 93% yield). M.p. = 156 °C (dec.). 1H NMR (500 MHz, DMSO-d6): δ 7.44 (d, J = 7.7 Hz, 2H), 7.37 (s, 1H), 7.08 (d, J = 7.7 Hz, 2H), 5.28 (s, 1H), 3.60–3.55 (m, 1H), 3.11–2.93 (m, 2H), 2.25 (s, 3H), 1.86–1.52 (dt, J = 85.6, 25.7 Hz, 22H) ppm. 13C{1H} NMR (126 MHz, DMSO-d6): δ 177.79, 151.60, 146.09, 144.80, 138.25, 128.63, 126.03, 110.33, 107.55, 87.84, 40.98, 36.73, 34.13, 33.08, 30.14, 28.29, 26.91, 25.06, 21.32 ppm.
(3-Amino-1H-pyrazol-1-yl) (dispiro[cyclohexane-1,3′-[1,2,4,5]tetraoxane-6′,2″-tricyclo [3.3.1.13,7] decan]-4-yl) methanone hydrochloride (T1•HCl). The compound was synthesised in accordance with general procedure 5, using T1. White solid (57 mg, 28% yield). 1H NMR (500 MHz, DMSO-d6): δ 7.41 (d, J = 1.8 Hz, 1H), 5.31 (d, J = 1.8 Hz, 1H), 3.61 (tt, J = 11.0, 3.7 Hz, 1H), 3.08 (d, J = 41.3 Hz, 2H), 1.86–1.64 (m, 22H) ppm. 13C{1H} NMR (126 MHz, DMSO-d6): δ 177.79, 151.60, 144.78, 110.33, 107.55, 87.85, 40.97, 36.73, 34.15, 33.08, 30.30, 29.98, 28.30, 26.91, 25.04, 24.08 ppm.
(3-Amino-1H-pyrazol-1-yl) (dispiro[cyclohexane-1,3′-[1,2,4,5]tetraoxane-6′,2″-tricyclo [3.3.1.13,7] decan]-4-yl) methanone hydrochloride (T2•HCl). The compound was synthesised in accordance with general procedure 5, using T2. White solid (23 mg, 11% yield). 1H NMR (500 MHz, DMSO-d6): δ 7.97 (d, J = 2.9 Hz, 1H), 5.88 (d, J = 3.0 Hz, 1H), 5.63 (s, 2H), 3.47–3.42 (m, 1H), 2.99 (br d, J = 48.0 Hz, 3H), 1.78–1.59 (m, 19H) ppm. 13C{1H} NMR (126 MHz, DMSO-d6): δ 176.34, 159.46, 129.89, 110.33, 110.25, 107.54, 102.56, 36.73, 34.15, 33.08, 31.77, 30.35, 29.54, 28.63, 26.91, 25.15, 24.44 ppm.

3.4. X-ray Crystallography

X-ray crystallographic studies were performed on single crystals of T1 and T2 at room temperature (292 K) for T2 and at low-temperature (150 K) for T1 and T2, using Mo Kα radiation (λ = 0.71073 Å). The room temperature data were collected in a Bruker Kappa Apex II diffractometer equipped with a 4K CCD detector; the low temperature data were measured in a Bruker D8 Quest diffractometer equipped with a PHOTON II CMOS detector and an Oxford Cryostream 700 N2 flow cryostat. Both isomers crystallize in the monoclinic P21/c space group (nº 14).
Summary of crystal data:
T1: a =14.909(4) Å, b= 11.695(3) Å, c = 11.121(3) Å, β = 108.697(7)°, V = 1836.7(3) Å 3, P21/c, Z = 4, T =150 (2) K, µ(Mo Kα) = 0.102 mm−1, Dcalc = 1.419 g/cm3, 8603 reflections measured (2.261° < 2θ < 30.187°) 3212 unique (Rint = 0.0542, Rσ = 0.0520), 2917 with I > 2σ which were used in the calculations. The final R1 was 0.1159 (I > 2σ) and wR was 0.3672 (all data).
T2: a =12.9853(2) Å, b= 11.2578(2) Å, c = 11.1712(2) Å, β = 101.911(6)°, V = 1884.0(4) Å3, P21/c, Z = 4, T =150 (2) K, µ(Mo Kα) = 0.099 mm−1, Dcalc = 1.373 g/cm3, 66967 reflections measured (2.402° < 2θ < 27.499°) 4321 unique (Rint = 0.0124, Rσ = 0.0322), 4030 with I > 2σ which were used in the calculations. The final R1 was 0.0449 (I > 2σ) and wR was 0.1139 (all data).
The structures of T1 and T2 were solved by direct methods using SHELXT-2018/2 [45]. Full-matrix least-squares refinements on F2 were carried out with SHELXL-2018/3 [46] with anisotropic displacement parameters for all non-hydrogen atoms (CCDC 2151742 and 2151175 for detailed crystal data).
Hydrogen atoms were located on a difference Fourier synthesis, and their positions were refined as riding on parent atoms with an isotropic temperature constrained those of their parent atoms using SHELXL-2018/3 defaults [46], except those of the amino group that had their positions freely refined with an isotropic displacement factor constrained to 1.2× that of the parent N atom, for being involved in hydrogen bonding.

3.5. Biological Evaluation

All compounds were solubilized in dimethyl sulfoxide (DMSO, Merck, Darmstadt, Germany). Working solutions for biological studies were prepared with a maximum of 1% DMSO. Amphotericin B (Gibco) was used as control drug.

3.5.1. Parasites and Cell Lines

For the anti-Leishmania assay, promastigote forms of L. infantum (MHOM/PT/1988/IMT151) and L. tropica (MHOM/PS/2002/63jnF21) were maintained at 24 °C ± 1 °C in M199 medium (Sigma), supplemented with 10% (v/v) foetal bovine serum (FBS, Sigma, Darmstadt, Germany), 2,5 mg/mL hemin (Sigma), HEPES 1M pH7.4 (Sigma), and 1% (v/v) penicillin-streptomycin (Sigma). The viscerotropic L. infantum strain belongs to IHMT, Portugal cryobank collection, and the dermotropic L. tropica strain was provided by the Technical University of Applied Sciences Wildau, Germany (Dr. Katrin Kuhls). Promastigotes in culture had less than 10 in vitro passages, aiming to keep the strain virulence. Promastigotes were used for the experiments at the end of the log phase (days 5–6).
THP-1 (ATCC® TIB-202TM) human acute monocytic leukemic cell line was obtained from DSMZ, Germany and cultured in RPMI 1640 (Sigma, Darmstadt, Germany) supplemented with 10% (v/v) FBS, 1% (v/v) penicillin-streptomycin (Sigma), and L-glutamine 200 mM (Sigma) at 37 °C, 5% CO2.

3.5.2. Leishmania spp. Promastigotes In Vitro Susceptibility

The half maximal inhibitory concentration (IC50) was determined. L. infantum and L. tropica promastigotes were platted in 96-well flat bottom culture plates (VWR) at a density of 5 × 106 parasites/mL in M199 medium with six different concentrations of each compound (OZ1: 400, 350, 300, 250, 200, and 150 µM; OZ2, T1, T2 and PIR: 400, 200, 100, 50, 25, and 12.5 µM), including amphotericin B (10, 5, 2.5, 1.25, 0.625, and 0.313 µM). A blank (M199 medium alone) and an untreated control (parasites with M199 medium plus 1% DMSO as vehicle solution) were also included. All the compounds were plated out in quadruplicate and incubated for 48-h at 24 °C ± 1 °C.
The parasite’s viability was determined using 3-(4,5-dimethyl-2-thiazolyl)-2,5-diphenyl-2H-tetrazolium bromide (MTT) colorimetric test. At the end of incubation time, 20 µL per well of MTT was added (5 mg/mL in PBS), followed by 2–4 h incubation at 37 °C ± 1 °C. After a centrifugation at 1000× g (20 min. 0 °C), medium was removed, and formazan crystals were solubilized in DMSO; optical density at 595 nm was measured with SynergyTM HTX Multi Mode Microplate Reader (Dynex Technologies, Chantilly, VA, USA). GraphPad Prism V 9.0 was used to calculate IC50 of each chemical by fitting the data as a non-linear regression with variable slope using a dose–response inhibitory model. At least three independent assays were performed.
To detect morphological alterations, at the end of the incubation period (48 h), 30 µL of the parasite’s suspension were collected from different treatment concentrations of each compound to prepare smears on glass slides, fixed with methanol, and stained with 5% (v/v) Giemsa (Sigma). Stained slides were examined with bright field in an inverted microscope (Eclipse 80i Nikon, Tokyo, Japan) with 1000× amplification, and images were captured with a Nikon DS-Ri1 camera. Image J (version 1.53k) was used to measure promastigote body and flagellum size of at least 30 parasite samples.
Data were expressed as mean ± standard error of the mean (SEM), and statistical analysis was performed with one-way ANOVA (Bonferoni’s multiple comparisons test) to determine whether differences between means relatively to untreated control were significant at different levels (p < 0.05, p < 0.01, p < 0.001).

3.5.3. In Vitro Cytotoxicity Assessment in THP-1 Cell Line

THP-1 cell line was used to calculate the endoperoxide–pyrazole hybrids’ half-maximal cytotoxic concentration (CC50). THP-1 was plated at a final density of 5 × 105 cells/mL, in 96-well flat bottom culture plates (VWR), using six different concentrations of the compounds (OZ1, OZ2, T1, T2 and PIR: 800, 400, 200, 100, 50, 25 µM) and amphotericin B (10, 5, 2.5, 1.25, 0.625, 0.313 µM). A blank (RPMI medium alone) and an untreated control (parasites with RPMI medium plus 1% DMSO as vehicle solution) were also included. All the compounds were plated out in quadruplicate and incubated for 48 h at 37 °C ± 1 °C, 5% CO2.
Following the incubation period, 20 µL per well of MTT was added as described in Section 3.5.2. After absorbance readings, the CC50 values of each compound were determined with GraphPad Prism V 9.0 (Dotmatics, San Diego, CA, USA) as previously described. At least three independent assays were performed.

3.5.4. Leishmania spp. Intracellular Amastigotes Susceptibility

The in vitro intracellular amastigote assay was performed using a THP1 cell line maintained in complete RPMI at 37 °C, 5% CO2. After 24 h differentiation of 5 × 105 cells/mL into macrophages in sterile 16-chamber LabTek slides (Thermo Scientific, Waltham, MA, USA) with 25 ng mL−1 phorbol myristic acid (PMA, Sigma), cells were washed once with warm RPMI to remove non-differentiated and non-adherent cells and further infected with 5 × 106 promastigotes/mL in a 10:1 parasite–macrophage ratio for another 24 h. After this period, slides were gently washed once with warm RPMI to remove non-internalised promastigotes. The compounds 2o, 2t, OZ1•HCl, OZ1•MsOH, and OZ1•TsOH at concentrations ranging 400 µM to 50 µM were added, and the slides were further incubated at 37 °C, 5% CO2 for 72 h. After this period, cells were washed with warm PBS, fixed with methanol (Sigma) and stained with Giemsa 5% (Sigma). AmB was used as positive control, and macrophages cultivated in RPMI medium and DMSO (0.5% v/v) were used as negative control. The infection index was determined by multiplying the percentage of infected cells and the number of amastigotes per infected cell as previously described [47]. The results represent the average of two experiments. The IC50 for each compound was calculated by fitting the data as a non-linear regression with variable slope using a dose–response inhibitory model in the GraphPad Prism V 8.0 (Dotmatics, San Diego, CA, USA) program.
Selectivity index (SI) was determined for each compound as the quotient between the cytotoxicity in cell lines (CC50) and the inhibitory concentration (IC50) (SI = CC50 in mammalian cells/IC50 L. infantum).

3.6. Software

Chemical structures were designed using ChemDraw Ultra 12.0 (Cambridge Soft, Cambridge, MA, USA). Graphics, measurements, and statistical analysis were performed using Microsoft Excel, ImageJ (version 1.53k, National Institutes of Health, Bethesda, Maryland, USA) and GraphPad Prism 9.0 (Dotmatics, San Diego, CA, USA). Bruker APEXIII [48] software package including SAINT V8.38A; [49] and SADABS-2016/2 [50] was used for data reduction of the X-ray crystallography work. SHELXT-2018/2 [45] and SHELXL-2018/3 [46] and PLATON (v. 2010420) [51] were used for the structure solution, refinement, and structure analysis and drawing of crystallographic figures, respectively.

4. Conclusions

This work describes, for the first time, the synthesis and structure of 1,2,4-trioxolane–pyrazole and 1,2,4,5-tetraoxane–pyrazole hybrids. Using a convergent synthetic approach, the hybrids were prepared by amide coupling of 3(5)-aminopyrazole with a carboxylic acid derivative of the endoperoxide-containing building block. The compounds were characterized in detail using 1D NMR (1H and 13C{1H}), 2D NMR (1H-1H COSY, HSQC, and HMBC), X-ray crystallography, and HRMS-ESI+. The 1H and 13C{1H} NMR spectroscopic and X-ray crystallography analyses enabled structural differentiation between 1,2,4-trioxolane and 1,2,4,5-tetraoxane isomers based on chemical shifts, signal sharpness, and number, whereas 1H-1H COSY and HMBC spectra allowed us to elucidate each positional isomer for both classes. X-ray crystallography unambiguously identified the structure of the two isomers and disclosed a distinct supramolecular organisation in the crystalline phase, despite both isomers crystallising in the same monoclinic space-group due to distinct hydrogen bonding patterns linking the molecules. Results reveal that only the endocyclic amine nitrogen of 3(5)-aminopyrazole participates in the amide coupling reactions with the carboxylic acid-containing endoperoxide building blocks. The reaction affords products from reaction of both 3- and 5-amino pyrazole tautomers, revealing the impact of prototropic tautomerism in the pyrazole heterocycle on reactivity.
Preliminary studies were conducted to evaluate the antileishmanial activity of this novel chemotype. Compounds OZ1 and OZ2 displayed better activity against L. tropica than the T1 and T2 compounds, which was attributed to the poor solubility of the tetraoxane hybrids in M199 and RPMI medium. Regarding the assessments performed in L. infantum, the results demonstrated no activity of all compounds toward this strain. The 3-aminopyrazole moiety was tested against both strains to determine the role of this moiety in the antileishmanial activity of all compounds and was found inactive. These results suggest that the antileishmanial activity depends on its substituents and could mean that the incorporation of a pyrazole scaffold into the endoperoxide structure may have no synergistic effect. In the future, and upon optimization, a broad range of Leishmania species and other protozoans, namely other trypanosomatids, should be tested in order to better understand their susceptibility and evaluate the potential of the novel chemotype.
The endoperoxide–pyrazole hybrids were found to be very stable, as their salt forms could be successfully prepared without compromising the endoperoxide pharmacophore. This conversion to the salt form solved the insoluble nature of these compounds. Tests were conducted to ascertain the potential activity of these endoperoxide–pyrazole hybrids in their salt forms against promastigote and intracellular amastigote forms. Among all the salts tested against promastigotes of L. tropica and L. infantum, OZ1•HCl proved to be the best option. Our findings indicate that the conversion of the free amine in the respective salts not only improved the solubility but also amplified the spectrum of action of this chemotype.
The evaluation of L. infantum amastigotes susceptibility corroborates the findings on the antileishmanial activity in promastigotes in which OZ1•HCl was the compound with better activity of the new series. In addition, the compound with the lowest IC50 against L. infantum amastigotes was identified as 2t, suggesting that the structure of these compounds must be simplified in order to enhance the effectiveness of new antileishmanial peroxides. Even though the issue of solubility has been addressed, it is not the primary factor influencing the antiparasitic effectiveness of this series. Considerable structure optimization was achieved with OZ1•HCl and 2t. However, the results of 1,2,4-trioxolane–pyrazole and 1,2,4,5-tetraoxane–pyrazole hybrids highlight the need for further structure optimization to improve the activity of these compounds.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules27175401/s1, S1. 1H, 13C{1H} NMR, COSY, HSQC, HMBC spectra of the synthesized compounds. S2. HRMS spectra of the synthesized compounds. S3. X-ray crystallography. Figure S1: 1H NMR spectrum (400 MHz) of 1o in CDCl3. Figure S2: 13C{1H} NMR spectrum (100 MHz) of 1o in CDCl3. Figure S3. 1H NMR spectrum (500 MHz) of 2o in CDCl3. Figure S4. 13C{1H} NMR spectrum (126 MHz) of 2o in CDCl3. Figure S5. 1H NMR spectrum (500 MHz) of 2t in CDCl3. Figure S6. 13C{1H} NMR spectrum (126 MHz) of 2t in CDCl3. Figure S7. 1H NMR spectrum (500 MHz) of 3o in CDCl3. Figure S8. 13C{1H} NMR spectrum (126 MHz) of 3o in CDCl3. Figure S9. 1H NMR spectrum (500 MHz) of 3t in CDCl3. Figure S10. 13C{1H} NMR spectrum (126 MHz) of 3t in CDCl3. Figure S11. 1H NMR spectrum (500 MHz) of OZ1 in CDCl3. Figure S12. 13C{1H} NMR spectrum (126 MHz) of OZ1 in CDCl3. Figure S13. 13C{1H} DEPT-135 NMR spectrum (126 MHz) of OZ1 in CDCl3. Figure S14. COSY spectrum (500 MHz) of OZ1 in CDCl3. Figure S15. HSQC spectrum (500 MHz) of OZ1 in CDCl3. Figure S16. HMBC spectrum (500 MHz) of OZ1 in CDCl3. Figure S17. 1H NMR spectrum (500 MHz) of OZ2 in CDCl3. Figure S18. 113C{1H} NMR spectrum (126 MHz) of OZ2 in CDCl3. Figure S19. COSY spectrum (500 MHz) of OZ2 in CDCl3. Figure S20. HSQC spectrum (500 MHz) of OZ2 in CDCl3. Figure S21. HMBC spectrum (500 MHz) of OZ2 in CDCl3. Figure S22. 1H NMR spectrum (500 MHz) of T1 in CDCl3. Figure S23. 13C{1H} NMR spectrum (126 MHz) of T1 in CDCl3. Figure S24. 13C{1H} DEPT-135 NMR spectrum (126 MHz) of T1 in CDCl3. Figure S25. COSY spectrum (500 MHz) of T1 in CDCl3. Figure S26. HSQC spectrum (500 MHz) of T1 in CDCl3. Figure S27. HMBC spectrum (500 MHz) of T1 in CDCl3. Figure S28. 1H NMR spectrum (500 MHz) of T2 in CDCl3. Figure S29. 13C{1H} NMR spectrum (126 MHz) of T2 in CDCl3. Figure S30. 13C{1H} DEPT-135 NMR spectrum (126 MHz) of T2 in CDCl3. Figure S31. COSY spectrum (500 MHz) of T2 in CDCl3. Figure S32. HSQC spectrum (500 MHz) of T2 in CDCl3. Figure S33. HMBC spectrum (500 MHz) of T2 in CDCl3. Figure S34. 1H NMR spectrum (500 MHz) of OZ1•TsOH in DMSO-d6. Figure S35. 13C{1H} NMR spectrum (126 MHz) of OZ1•TsOH in DMSO-d6. Figure S36. 1H NMR spectrum (500 MHz) of OZ1•MsOH in DMSO-d6. Figure S37. 13C{1H} NMR spectrum (126 MHz) of OZ1•MsOH in DMSO-d6. Figure S38. 1H NMR spectrum (500 MHz) of OZ2•TsOH in DMSO-d6. Figure S39. 13C{1H} NMR spectrum (126 MHz) of OZ2•TsOH in DMSO-d6. Figure S40. 1H NMR spectrum (500 MHz) of T1•TsOH in DMSO-d6. Figure S41. 13C{1H} NMR spectrum (126 MHz) of T1•TsOH in DMSO-d6. Figure S42. 1H NMR spectrum (500 MHz) of OZ1•HCl in DMSO-d6. Figure S43. 13C{1H} NMR spectrum (126 MHz) of OZ1•HCl in DMSO-d6. Figure S44. 1H NMR spectrum (500 MHz) of OZ2•HCl in DMSO-d6. Figure S45. 13C{1H} NMR spectrum (126 MHz) of OZ2•HCl in DMSO-d6. Figure S46. 1H NMR spectrum (500 MHz) of T1•HCl in DMSO-d6. Figure S47. 13C{1H} NMR spectrum (126 MHz) of T1•HCl in DMSO-d6. Figure S48. 1H NMR spectrum (500 MHz) of T2•HCl in DMSO-d6. Figure S49. 13C{1H} NMR spectrum (126 MHz) of T2•HCl in DMSO-d6. Figure S50. ORTEP drawing of the molecule of (5-amino-1H-pyrazol-1-yl) (dispiro[cyclohexane-1,3′-[1,2,4,5]tetraoxane-6′,2″-tricyclo[3.3.1.13,7]decan]-4-yl)-methanone (T1). Figure S51. ORTEP drawing of the molecule of (3-amino-1H-pyrazol-1-yl) (dispiro[cyclohexane-1,3′-[1,2,4,5]tetraoxane-6′,2″-tricyclo[3.3.1.13,7]-decan]-4-yl)-methanone (T2).

Author Contributions

Conceptualization, P.S.M.A. and M.L.S.C.; methodology (synthesis and structure analysis): P.S.M.A.; methodology (in vitro tests): I.C.C.C. and S.C; methodology (X-ray crystallography): R.F.M. and J.A.P.; validation, all authors; formal analysis, all authors; investigation, all authors; resources, S.C., J.A.P. and M.L.S.C.; writing—original draft preparation, P.S.M.A., I.C.C.C. and J.A.P.; writing—review and editing, S.C. and M.L.S.C.; visualization, P.S.M.A., I.C.C.C., J.A.P.; supervision, M.L.S.C.; project administration, M.L.S.C.; funding acquisition, S.C., J.A.P. and M.L.S.C. All authors have read and agreed to the published version of the manuscript.

Funding

This study received Portuguese national funds from FCT—the Foundation for Science and Technology, through projects UIDB/04326/2020, UIDP/04326/2020 and LA/P/0101/2020 (CCMar) UIDB/04564/2020, and UIDP/04564/2020 (CFisUC), IF/00743/2015/CP1320, UID/Multi/04413/2019 (GHTM), UIDB/50011/2020, UIDP/50011/2020 & LA/P/0006/2020, financed by national funds through the FCT/MEC (PIDDAC), and from the operational programmes CRESC Algarve 2020 and COMPETE 2020 through project EMBRC.PT ALG-01-0145-FEDER-022121. We gratefully acknowledge FCT for Grants SFRH/BD/130407/2017 and COVID/BD/152392/2022 (P.S.M.A.), SFRH/BD/08242/2020 (I.C.C.C.) and for the Junior Research Position CEECIND/00553/2017 (R.F.M.).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

CIF files containing supplementary crystallographic data were deposited at the Cambridge Crystallographic Data Centre with references 2151742 (T1) and 2151175 (T2).

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

The samples are available from the authors.

References

  1. Cabral, L.I.L.; Pomel, S.; Cojean, S.; Amado, P.S.M.; Loiseau, P.M.; Cristiano, M.L.S. Synthesis and Antileishmanial Activity of 1,2,4,5-Tetraoxanes against Leishmania Donovani. Molecules 2020, 25, 465. [Google Scholar] [CrossRef]
  2. Mann, S.; Frasca, K.; Scherrer, S.; Henao-Martínez, A.F.; Newman, S.; Ramanan, P.; Suarez, J.A. A Review of Leishmaniasis: Current Knowledge and Future Directions. Curr. Trop. Med. Reports 2021, 8, 121–132. [Google Scholar] [CrossRef] [PubMed]
  3. Roatt, B.M.; de Oliveira Cardoso, J.M.; De Brito, R.C.F.; Coura-Vital, W.; de Oliveira Aguiar-Soares, R.D.; Reis, A.B. Recent Advances and New Strategies on Leishmaniasis Treatment. Appl. Microbiol. Biotechnol. 2020, 104, 8965–8977. [Google Scholar] [CrossRef] [PubMed]
  4. Azim, M.; Khan, S.A.; Ullah, S.; Ullah, S.; Anjum, S.I. Therapeutic Advances in the Topical Treatment of Cutaneous Leishmaniasis: A Review. PLoS Negl. Trop. Dis. 2021, 15, e0009099. [Google Scholar] [CrossRef] [PubMed]
  5. Cortes, S.; Albuquerque, A.; Cabral, L.I.L.; Lopes, L.; Campino, L.; Cristiano, M.L.S. In Vitro Susceptibility of Leishmania Infantum to Artemisinin Derivatives and Selected Trioxolanes. Antimicrob. Agents Chemother. 2015, 59, 5032–5035. [Google Scholar] [CrossRef] [PubMed]
  6. Antolínez, I.V.; Barbosa, L.C.A.; Borgati, T.F.; Baldaia, A.; Ferreira, S.R.; Almeida, R.M.; Fujiwara, R.T. Tetraoxanes as New Agents Agains Leishmania Amazonensis. Chem. Biodivers. 2020, 17, e2000142. [Google Scholar] [CrossRef] [PubMed]
  7. Medrán, N.S.; Sayé, M.; Pereira, C.A.; Tekwani, B.L.; La-Venia, A.; Labadie, G.R. Expanding the Scope of Synthetic 1,2,4-Trioxanes towards Trypanosoma Cruzi and Leishmania Donovani. Bioorg. Med. Chem. Lett. 2020, 30, 127491. [Google Scholar] [CrossRef]
  8. Mendes, A.; Armada, A.; Cabral, L.I.L.; Amado, P.S.M.; Campino, L.; Cristiano, M.L.S.; Cortes, S. 1,2,4-Trioxolane and 1,2,4,5-Tetraoxane Endoperoxides against Old-World Leishmania Parasites: In Vitro Activity and Mode of Action. Pharmaceuticals 2022, 15, 446. [Google Scholar] [CrossRef]
  9. Opsenica, D.M.; Šolaja, B.A. Antimalarial Peroxides. J. Serbian Chem. Soc. 2009, 74, 1155–1193. [Google Scholar] [CrossRef]
  10. Amewu, R.K.; Chadwick, J.; Hussain, A.; Panda, S.; Rinki, R.; Janneh, O.; Ward, S.A.; Miguel, C.; Burrell-Saward, H.; Vivas, L.; et al. Synthesis and Evaluation of the Antimalarial, Anticancer, and Caspase 3 Activities of Tetraoxane Dimers. Bioorganic Med. Chem. 2013, 21, 7392–7397. [Google Scholar] [CrossRef]
  11. Woodley, C.M.; Amado, P.S.M.; Cristiano, M.L.S.; O’Neill, P.M. Artemisinin Inspired Synthetic Endoperoxide Drug Candidates: Design, Synthesis, and Mechanism of Action Studies. Med. Res. Rev. 2021, 41, 3062–3095. [Google Scholar] [CrossRef] [PubMed]
  12. O’Neill, P.M.; Amewu, R.K.; Nixon, G.L.; ElGarah, F.B.; Mungthin, M.; Chadwick, J.; Shone, A.E.; Vivas, L.; Lander, H.; Barton, V.; et al. Identification of a 1,2,4,5-Tetraoxane Antimalarial Drug-Development Candidate (RKA 182) with Superior Properties to the Semisynthetic Artemisinins. Angew. Chem.-Int. Ed. 2010, 49, 5693–5697. [Google Scholar] [CrossRef]
  13. O’Neill, P.M.; Sabbani, S.; Nixon, G.L.; Schnaderbeck, M.; Roberts, N.L.; Shore, E.R.; Riley, C.; Murphy, B.; Mcgillan, P.; Ward, S.A.; et al. Optimisation of the synthesis of second generation 1,2,4,5 tetraoxane antimalarials. Tetrahedron 2016, 72, 6118–6126. [Google Scholar] [CrossRef]
  14. Mohamed, L.W.; Shaaban, M.A.; Zaher, A.F.; Alhamaky, S.M.; Elsahar, A.M. Synthesis of New Pyrazoles and Pyrozolo [3,4-b] Pyridines as Anti-Inflammatory Agents by Inhibition of COX-2 Enzyme. Bioorg. Chem. 2019, 83, 47–54. [Google Scholar] [CrossRef] [PubMed]
  15. Bennani, F.E.; Doudach, L.; Cherrah, Y.; Ramli, Y.; Karrouchi, K.; Ansar, M.; Faouzi, M.E.A. Overview of Recent Developments of Pyrazole Derivatives as an Anticancer Agent in Different Cell Line. Bioorg. Chem. 2020, 97, 103470. [Google Scholar] [CrossRef] [PubMed]
  16. Ouyang, G.; Cai, X.-J.; Chen, Z.; Song, B.-A.; Bhadury, P.S.; Yang, S.; Jin, L.-H.; Xue, W.; Hu, D.-Y.; Zeng, S. Synthesis and Antiviral Activities of Pyrazole Derivatives Containing an Oxime Moiety. J. Agric. Food Chem. 2008, 56, 10160–10167. [Google Scholar] [CrossRef] [PubMed]
  17. Bekhit, A.A.; Saudi, M.N.; Hassan, A.M.M.; Fahmy, S.M.; Ibrahim, T.M.; Ghareeb, D.; El-Seidy, A.M.; Al-Qallaf, S.M.; Habib, H.J.; Bekhit, A.E.D.A. Synthesis, Molecular Modeling and Biological Screening of Some Pyrazole Derivatives as Antileishmanial Agents. Future Med. Chem. 2018, 10, 2325–2344. [Google Scholar] [CrossRef]
  18. Secrieru, A.; O’Neill, P.M.; Cristiano, M.L.S. Revisiting the Structure and Chemistry of 3(5)-Substituted Pyrazoles. Molecules 2020, 25, 42. [Google Scholar] [CrossRef]
  19. Castillo, J.-C.; Portilla, J. Recent Advances in the Synthesis of New Pyrazole Derivatives. Targets Heterocycl. Syst. 2018, 22, 194–223. [Google Scholar] [CrossRef]
  20. Gao, C.; Chang, L.; Xu, Z.; Yan, X.-F.; Ding, C.; Zhao, F.; Wu, X.; Feng, L.S. Recent Advances of Tetrazole Derivatives as Potential Anti-Tubercular and Anti-Malarial Agents. Eur. J. Med. Chem. 2019, 163, 404–412. [Google Scholar] [CrossRef]
  21. Uddin, A.; Chawla, M.; Irfan, I.; Mahajan, S.; Singh, S.; Abid, M. Medicinal Chemistry Updates on Quinoline- and Endoperoxide-Based Hybrids with Potent Antimalarial Activity. RSC Med. Chem. 2021, 12, 24–42. [Google Scholar] [CrossRef] [PubMed]
  22. Sharma, B.; Singh, P.; Singh, A.K.; Awasthi, S.K. Advancement of Chimeric Hybrid Drugs to Cure Malaria Infection: An Overview with Special Emphasis on Endoperoxide Pharmacophores. Eur. J. Med. Chem. 2021, 219, 113408. [Google Scholar] [CrossRef] [PubMed]
  23. Tibon, N.S.; Ng, C.H.; Cheong, S.L. Current Progress in Antimalarial Pharmacotherapy and Multi-Target Drug Discovery. Eur. J. Med. Chem. 2020, 188, 111983. [Google Scholar] [CrossRef] [PubMed]
  24. Campino, L.; Pratlong, F.; Abranches, P.; Rioux, J.-A.; Santos-Gomes, G.; Alves-Pires, C.; Cortes, S.; Ramada, J.; Cristovão, J.M.; Afonso, M.O.; et al. Leishmaniasis in Portugal: Enzyme Polymorphism of Leishmania Infantum Based on the Identification of 213 Strains. Trop. Med. Int. Health 2006, 11, 1708–1714. [Google Scholar] [CrossRef] [PubMed]
  25. Saik, I.E.I.; Benlabsir, C.; Fellah, H.; Lemrani, M.; Riyad, M. Transmission Patterns of Leishmania Tropica around the Mediterranean Basin: Could Morocco Be Impacted by a Zoonotic Spillover? PLoS Negl. Trop. Dis. 2022, 16, e0010009. [Google Scholar] [CrossRef]
  26. Amado, P.S.M.; Frija, L.M.T.; Coelho, J.A.S.; O’Neill, P.M.; Cristiano, M.L.S. Synthesis of Non-Symmetrical Dispiro-1,2,4,5-Tetraoxanes and Dispiro-1,2,4-Trioxanes Catalyzed by Silica Sulfuric Acid. J. Org. Chem. 2021, 86, 10608–10620. [Google Scholar] [CrossRef] [PubMed]
  27. Dong, Y.; Vennerstrom, J.L. Differentiation between 1,2,4,5-Tetraoxanes and 1,2,4,5,7,8- Hexaoxonanes Using 1H and 13C NMR Analyses. J. Heterocycl. Chem. 2001, 38, 463–466. [Google Scholar] [CrossRef]
  28. Secrieru, A.; Lopes, S.; Cristiano, M.L.S.; Fausto, R. UV-Induced Tautomerization in Argon Matrix. Molecules 2021, 26, 4299. [Google Scholar] [CrossRef] [PubMed]
  29. Spek, A.L. Single-crystal structure validation with the program PLATON. J. Appl. Crystallogr. 2003, 36, 7–13. [Google Scholar] [CrossRef]
  30. Cremer, D.; Pople, J.A. A General Definition of Ring Puckering Coordinates. J. Am. Chem. Soc. 1975, 97, 1354–1358. [Google Scholar] [CrossRef]
  31. Cortes, S.; Sousa, C.B.; Morais, T.; Lago, J.; Campino, L. Potential of the natural products against leishmaniasis in Old World-a review of in-vitro studies. Pathog. Glob. Health 2020, 114, 170–182. [Google Scholar] [CrossRef] [PubMed]
  32. Madusanka, R.K.; Silva, H.; Karunaweera, N.D. Treatment of Cutaneous Leishmaniasis and Insights into Species-Specific Responses: A Narrative Review. Infect. Dis. Ther. 2022, 11, 695–711. [Google Scholar] [CrossRef] [PubMed]
  33. Maurício, I.L. Leishmania Taxonomy. In The Leishmaniases: Old Neglected Tropical Diseases; Springer: Cham, Switzerland, 2018; pp. 15–30. [Google Scholar] [CrossRef]
  34. Machín, L.; Nápoles, R.; Gille, L.; Monzote, L. Leishmania Amazonensis Response to Artemisinin and Derivatives. Parasitol. Int. 2021, 80, 102218. [Google Scholar] [CrossRef] [PubMed]
  35. Woodley, C.M.; Nixon, G.L.; Basilico, N.; Parapini, S.; Hong, W.D.; Ward, S.A.; Biagini, G.A.; Leung, S.C.; Taramelli, D.; Onuma, K.; et al. Enantioselective Synthesis and Profiling of Potent, Nonlinear Analogues of Antimalarial Tetraoxanes E209 and N205. ACS Med. Chem. Lett. 2021, 12, 1077–1085. [Google Scholar] [CrossRef]
  36. Brown, R.J.C.; Brown, R.F.C. Melting Point and Molecular Symmetry. J. Chem. Educ. 2000, 77, 724. [Google Scholar] [CrossRef]
  37. Gautam, A.; Ahmed, T.; Paliwal, J.; Batra, V. Pharmacokinetics and Pharmacodynamics of Endoperoxide Antimalarials. Curr. Drug Metab. 2009, 10, 289–306. [Google Scholar] [CrossRef]
  38. Adebayo, J.O.; Tijjani, H.; Adegunloye, A.P.; Ishola, A.A.; Balogun, E.A.; Malomo, S.O. Enhancing the antimalarial activity of artesunate. Parasitol. Res. 2020, 119, 2749–2764. [Google Scholar] [CrossRef]
  39. Bruno de Sousa, C.; Lago, J.H.G.; Macridachis, J.; Oliveira, M.; Brito, L.; Vizetto-Duarte, C.; Florindo, C.; Hendrickx, S.; Maes, L.; Morais, T.; et al. Report of in Vitro Antileishmanial Properties of Iberian Macroalgae. Nat. Prod. Res. 2019, 33, 1778–1782. [Google Scholar] [CrossRef]
  40. Hendrickx, S.; Guerin, P.J.; Caljon, G.; Croft, S.L.; Maes, L. Evaluating drug resistance in visceral leishmaniasis: The challenges. Parasitology 2018, 145, 453–463. [Google Scholar] [CrossRef]
  41. Jiricek, J.; Kondreddi, R.R.; Smith, P.W. Indole Carboxamide Derivatives and Uses Thereof. International Patent WO2014037900A1, March 2014. [Google Scholar]
  42. Stec, J.; Onajole, O.K.; Lun, S.; Guo, H.; Merenbloom, B.; Vistoli, G.; Bishai, W.R.; Kozikowski, A.P. Indole-2-Carboxamide-Based MmpL3 Inhibitors Show Exceptional Antitubercular Activity in an Animal Model of Tuberculosis Infection. J. Med. Chem. 2016, 59, 6232–6247. [Google Scholar] [CrossRef]
  43. Vennerstrom, J.L.; Dong, Y.; Chollet, J.; Matile, H.; Wang, X.; Spiraghavan, K.; Chapman, W.N. Spiro and Dispiro 1,2,4-Trioxolane Antimalarials. U.S. Patent US 7 371 778 B2, May 2008. [Google Scholar]
  44. Lis, L.; Cabral, L.I.L.; Sena, M.I.; Guerreiro, B.; Rodrigues, A.S.; De Andrade-Neto, V.F.; Cristiano, M.L.S.; Nogueira, F. New endoperoxides highly active in vivo and in vitro against artemisinin-resistant Plasmodium Falciparum. Malar. J. 2018, 17, 145. [Google Scholar] [CrossRef]
  45. Sheldrick, G.M. SHELXT—Integrated Space-Group and Crystal- Structure Determination. Acta Crystallogr. Sect. A 2015, A71, 3–8. [Google Scholar] [CrossRef] [PubMed]
  46. Sheldrick, G.M. Crystal Structure Refinement with SHELXL. Acta Crystallogr. Sect. C 2015, C71, 3–8. [Google Scholar] [CrossRef]
  47. Yamamoto, E.S.; Campos, B.L.S.; Jesus, J.A.; Laurenti, M.D.; Ribeiro, S.P.; Kallás, E.G.; Rafael-Fernandes, M.; Santos-Gomes, G.; Silva, M.S.; Sessa, D.P.; et al. The Effect of Ursolic Acid on Leishmania (Leishmania) amazonensis Is Related to Programed Cell Death and Presents Therapeutic Potential in Experimental Cutaneous Leishmaniasis. PLoS ONE 2015, 10, e0144946. [Google Scholar] [CrossRef] [PubMed]
  48. Bruker. APEX3; Bruker AXS Inc.: Madison, WI, USA, 2016. [Google Scholar]
  49. Bruker. SAINT V8.38A; Bruker AXS Inc.: Madison, WI, USA, 2016. [Google Scholar]
  50. Krause, L.; Herbst-Irmer, R.; Sheldrick, G.M.; Stalke, D. Comparison of silver and molybdenum microfocus X-ray sources for single-crystal structure determination. J. Appl. Cryst. 2015, 48, 3–10. [Google Scholar] [CrossRef]
  51. Spek, A.L. Structure validation in chemical crystallography. Acta Cryst. D 2009, D65, 148–155. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Structures of: (A) 1,2,4-trioxolane LC95 [5], 1,2,4,5-tetraoxane LC140 [1], and pyrazole derivative 3i [17], with known antileishmanial activity; and (B) 1,2,4-trioxolane-3(5)-aminopyrazoles (OZ1, OZ2) and 1,2,4,5-tetraoxane-3(5)-aminopyrazoles (T1, T2) reported in this study.
Figure 1. Structures of: (A) 1,2,4-trioxolane LC95 [5], 1,2,4,5-tetraoxane LC140 [1], and pyrazole derivative 3i [17], with known antileishmanial activity; and (B) 1,2,4-trioxolane-3(5)-aminopyrazoles (OZ1, OZ2) and 1,2,4,5-tetraoxane-3(5)-aminopyrazoles (T1, T2) reported in this study.
Molecules 27 05401 g001
Scheme 1. Reagents and conditions: (a) pyridine, MeONH2, MeOH, r.t, 72 h; (b) ethyl 4-oxocyclohexanecarboxylate, O3, DCM/Pentane, 0 °C, 2 h; (c) LiOH, THF/H2O, rt, o/n; (d) EDC.HCl, HOBt, Et3N, CH2Cl2, rt, o/n.
Scheme 1. Reagents and conditions: (a) pyridine, MeONH2, MeOH, r.t, 72 h; (b) ethyl 4-oxocyclohexanecarboxylate, O3, DCM/Pentane, 0 °C, 2 h; (c) LiOH, THF/H2O, rt, o/n; (d) EDC.HCl, HOBt, Et3N, CH2Cl2, rt, o/n.
Molecules 27 05401 sch001
Scheme 2. Reagents and conditions: (a) H2SO4-SiO2, H2O2 (50% w/w), CH3CN, 0º C-rt, 1 h; (b) adamantan-2-one, H2SO4-SiO2, CH2Cl2, 0ª C-r.t., 3 h; (c) LiOH, THF/H2O, rt, o/n; (d) EDC.HCl, HOBt, Et3N, CH2Cl2, rt, o/n.
Scheme 2. Reagents and conditions: (a) H2SO4-SiO2, H2O2 (50% w/w), CH3CN, 0º C-rt, 1 h; (b) adamantan-2-one, H2SO4-SiO2, CH2Cl2, 0ª C-r.t., 3 h; (c) LiOH, THF/H2O, rt, o/n; (d) EDC.HCl, HOBt, Et3N, CH2Cl2, rt, o/n.
Molecules 27 05401 sch002
Figure 2. ORTEP drawing of the molecule of (5-amino-1H-pyrazol-1-yl) (dispiro[cyclohexane-1,3′-[1,2,4,5]tetraoxane-6′,2″-tricyclo[3.3.1.13,7]decan]-4-yl)-methanone (T1). Displacement ellipsoids are drawn at the 50% probability level.
Figure 2. ORTEP drawing of the molecule of (5-amino-1H-pyrazol-1-yl) (dispiro[cyclohexane-1,3′-[1,2,4,5]tetraoxane-6′,2″-tricyclo[3.3.1.13,7]decan]-4-yl)-methanone (T1). Displacement ellipsoids are drawn at the 50% probability level.
Molecules 27 05401 g002
Figure 3. ORTEP drawing of the molecule of (3-amino-1H-pyrazol-1-yl)-(dispiro[cyclohexane-1,3′-[1,2,4,5]tetraoxane-6′,2″-tricyclo[3.3.1.13,7]decan]-4-yl)-methanone (T2). Displacement ellipsoids are drawn at the 50% probability level. The tetraoxane group has minor disorder over two alternate chair conformations; for the sake of clarity, only the major conformation is shown.
Figure 3. ORTEP drawing of the molecule of (3-amino-1H-pyrazol-1-yl)-(dispiro[cyclohexane-1,3′-[1,2,4,5]tetraoxane-6′,2″-tricyclo[3.3.1.13,7]decan]-4-yl)-methanone (T2). Displacement ellipsoids are drawn at the 50% probability level. The tetraoxane group has minor disorder over two alternate chair conformations; for the sake of clarity, only the major conformation is shown.
Molecules 27 05401 g003
Figure 4. Hydrogen-bonding motif in the crystal structure of isomer T1.
Figure 4. Hydrogen-bonding motif in the crystal structure of isomer T1.
Molecules 27 05401 g004
Figure 5. Hydrogen-bonding motif in the crystal structure of isomer T2 (projection along the crystallographic a-axis).
Figure 5. Hydrogen-bonding motif in the crystal structure of isomer T2 (projection along the crystallographic a-axis).
Molecules 27 05401 g005
Figure 6. Body (A) and flagellum (B) measurements of L. tropica promastigotes after 48 h of treatment with OZ1 (300 µM), OZ2 (200 µM), and amphotericin B (0.62 µM). Non-treated parasites were also included in this evaluation. Significant differences between compounds and amphotericin B in comparison with non-treated parasites are indicated as *, p < 0.001 (ANOVA).
Figure 6. Body (A) and flagellum (B) measurements of L. tropica promastigotes after 48 h of treatment with OZ1 (300 µM), OZ2 (200 µM), and amphotericin B (0.62 µM). Non-treated parasites were also included in this evaluation. Significant differences between compounds and amphotericin B in comparison with non-treated parasites are indicated as *, p < 0.001 (ANOVA).
Molecules 27 05401 g006
Figure 7. Assessment of L. tropica promastigotes morphology after 48- hours of treatment with 300 µM of OZ1 (C,D), 200 µM of OZ2 (E,F), and 0.625 µM of amphotericin B (B); in comparison with non-treated parasites (A). Figures (CF) show the damage caused by both ozonides to the parasite, including body size reduction with an unnaturally rounded form, cytoplasmic disaggregation, derangement of internal organelles, and cell shrinkage. The observations were conducted on an inverted microscope (Eclipse 80i Nikon), and images were captured with a Nikon DS-Ri1 camera (magnification ×100).
Figure 7. Assessment of L. tropica promastigotes morphology after 48- hours of treatment with 300 µM of OZ1 (C,D), 200 µM of OZ2 (E,F), and 0.625 µM of amphotericin B (B); in comparison with non-treated parasites (A). Figures (CF) show the damage caused by both ozonides to the parasite, including body size reduction with an unnaturally rounded form, cytoplasmic disaggregation, derangement of internal organelles, and cell shrinkage. The observations were conducted on an inverted microscope (Eclipse 80i Nikon), and images were captured with a Nikon DS-Ri1 camera (magnification ×100).
Molecules 27 05401 g007
Scheme 3. Preparation of different salt forms from the endoperoxide–pyrazole hybrids.
Scheme 3. Preparation of different salt forms from the endoperoxide–pyrazole hybrids.
Molecules 27 05401 sch003
Figure 8. Graphic representation of cytotoxic (CC50) and inhibitory concentrations (IC50) for THP1 cell line, promastigotes, and amastigotes of L. infantum, respectively.
Figure 8. Graphic representation of cytotoxic (CC50) and inhibitory concentrations (IC50) for THP1 cell line, promastigotes, and amastigotes of L. infantum, respectively.
Molecules 27 05401 g008
Table 1. Correlations found in the HSQC, and HMBC spectra of 1,2,4-trioxolane–pyrazole isomers OZ1 and OZ2.
Table 1. Correlations found in the HSQC, and HMBC spectra of 1,2,4-trioxolane–pyrazole isomers OZ1 and OZ2.
Molecules 27 05401 i001
Isomer OZ1Isomer OZ2
1H, δ (ppm)13C, δ (ppm)1H, δ (ppm)13C, δ (ppm)
COSYHSQCHMBC COSYHSQCHMBC
1.671.9736.8526.75, 35.05, 111.731.671.72, 1.80, 1.8130.20, 36.8526.75, 34.95, 111.49, 111.73
1.691.9734.85111.731.681.7334.8526.79, 111.49
1.722.00, 1.9734.8526.751.721.6735.82-
1.792.01, 2.02, 3.5726.5526.501.731.67, 1.68, 2.0234.85-
1.813.5733.5236.76, 108.071.751.97, 2.0233.69-
1.823.5727.05, 26.2233.32, 33.10, 40.911.771.67, 1.90, 1.9626.5526.53, 39.69
1.971.67, 1.69, 1.72, 3.5736.5136.76, 36.50, 111.731.801.67, 1.9033.3636.89, 39.70, 108.07
2.001.72, 3.5734.8534.811.811.67, 1.9026.3839.73
2.011.79, 3.5726.38 1.821.7227.0526.50, 39.73
2.021.79, 3.5733.5226.26, 33.34, 40.911.901.8134.8526.58, 33.84, 37.02
3.571.79, 1.81, 1.82, 1.97, 2.00, 2.0140.83-1.961.67, 1.7726.38. 33.5239.8, 108.07, 173.12
5.19--NH89.021.971.75, 1.77, 1.8134.85, 36.5126.53, 36.85, 111.49, 173.12
5.375.54, 7.3688.02143.97,150.322.021.73, 1.7533.5226.50, 33.40, 39.69, 108.07
5.545.37, 7.36-NH89.023.98--NH2100.99, 157.40
7.365.37, 5.54143.9789.02, 150.325.887.98100.99130.18, 157.40
----7.985.88130.18100.99; 157.40
Table 2. Correlations found in the HSQC, and HMBC spectra of 1,2,4,5-tetraoxane–pyrazole isomers T1 and T2.
Table 2. Correlations found in the HSQC, and HMBC spectra of 1,2,4,5-tetraoxane–pyrazole isomers T1 and T2.
Molecules 27 05401 i002
Isomer T1Isomer T2
1H, δ (ppm)13C, δ (ppm)1H, δ (ppm)13C, δ (ppm)
COSYHSQCHMBC COSYHSQCHMBC
1.601.62, 1.70, 1.73, 1.8633.19-1.591.6933.19-
1.621.60, 1.70, 1.73,30.47-1.691.59, 1.7137.0133.34
1.701.7337.0133.341.713.14, 1.7637.0126.99, 33.34
1.731.70, 3.1633.1926.991.761.69, 1.71, 3.14, 1.9733.19-
1.861.97, 3.16, 3.6530.56, 33.19-1.833.4827.21-
1.971.73, 3.6527.05-1.871.76, 2.15, 3.4828.2130.17, 34.07
3.161.7330.2-1.971.76, 2.1533.19172.78
3.651.86, 1.9741.33178.12.021.87. 1.9733.19-
5.19--NH88.923.141.71, 1.76--
5.377.36, 5.5488.92143.97, 150.323.481.85, 1.8740.00172.78
5.545.37, 7.36-NH88.923.978.00-NH2101.04, 157.40
7.365.37; 5.54144.0489.02, 150.325.888.00101.04130.14, 157.40
----8.003.97, 5.88130.14101.04, 157.40
Table 3. Hydrogen bonding information for the crystalline structures of isomers T1 and T2. Symmetry codes: a: x,y,−1+z; b: 1−x,2−y,1−z; c: 1−x,1/2+y,1/2−z.
Table 3. Hydrogen bonding information for the crystalline structures of isomers T1 and T2. Symmetry codes: a: x,y,−1+z; b: 1−x,2−y,1−z; c: 1−x,1/2+y,1/2−z.
T1: D–HAD–H (Å)HA (Å)DA (Å)<D–HA (°)
N3–H3AN2 a0.882.553.528 (11)137
N3–H3BO50.882.152.755 (11)125
T2: D–HAD–H (Å)HA (Å)DA (Å)<D–HA (°)
N3–H3AN2 b0.868 (18)2.192 (18)3.0432 (16)166.7 (17)
N3–H3BO5 c0.830 (18)2.513 (18)3.1726 (17)137.2 (16)
Table 4. Estimated ClogP and ClogS values, cytotoxic concentrations (CC50) in THP-1 cell line, inhibitory concentrations (IC50) in promastigotes of L. infantum and L. tropica and intracellular amastigotes of L. infantum, and selectivity index (SI) of synthetic endoperoxide–pyrazole hybrids, and amphotericin B (control).
Table 4. Estimated ClogP and ClogS values, cytotoxic concentrations (CC50) in THP-1 cell line, inhibitory concentrations (IC50) in promastigotes of L. infantum and L. tropica and intracellular amastigotes of L. infantum, and selectivity index (SI) of synthetic endoperoxide–pyrazole hybrids, and amphotericin B (control).
IDStructuresClogP aClogS aCytotoxicity CC50 ± SEM b
(µΜ)
Promastigotes
Susceptibility
IC50 ± SEM b (µΜ)
Intracellular Amastigote Susceptibility
IC50 ± SEM c (µΜ)
SI d
L. tropicaL. infantumL. infantum
2oMolecules 27 05401 i0032.611.38381 ± 172156 ± 18123 ± 1330 ± 481.2
2tMolecules 27 05401 i0042.381.21374 ± 119143 ± 9128 ± 2745 ± 288.4
OZ1Molecules 27 05401 i0052.011.30664 ± 163219 ± 25>400NDND
OZ2Molecules 27 05401 i0062.011.30202 ± 116161 ± 19219 ± 44NDND
T1Molecules 27 05401 i0071.791.18>800>400>400NDND
T2Molecules 27 05401 i0081.791.18>800NDNDNDND
PIRMolecules 27 05401 i009−0.515.59202 ± 10>400>400NDND
AmBAmphotericin B−1.192.4823 ± 30.30 ± 0.10.20 ± 0.010.08 ± 0.05290
a Calculated using StarDrop v7.2.0.32905; b SEM: standard error of mean of at least three independent assays; c SEM: standard error of mean of two independent assays; d SI for amastigote forms; ND: not defined.
Table 5. Cytotoxic concentrations (CC50) in THP-1 cell line, inhibitory concentrations (IC50) in promastigote forms of L. infantum and L. tropica, and in intracellular amastigote forms of L. infantum and selective index (SI) of synthetic endoperoxide–pyrazole salts.
Table 5. Cytotoxic concentrations (CC50) in THP-1 cell line, inhibitory concentrations (IC50) in promastigote forms of L. infantum and L. tropica, and in intracellular amastigote forms of L. infantum and selective index (SI) of synthetic endoperoxide–pyrazole salts.
CompoundsPromastigotes Susceptibility IC50 ± SEM a (µΜ)Cytotoxicity
CC50 ± SEM a (µΜ)
Intracellular Amastigotes Susceptibility IC50 ± SEM b (µΜ)SI c
L. tropicaL. infantumL. infantum
OZ1•TsOH199 ± 9181 ± 35510 ± 109246 ± 322.1
OZ1•MsOH198 ± 24183 ± 48336 ± 196>400<0.8
OZ1•HCl135 ± 36164 ± 20509 ± 15587 ± 395.9
OZ2•TsOH260 ± 82217 ± 47479 ± 83NDND
OZ2•HCl>400258 ± 91425 ± 104NDND
T1•TsOH>400>400331 ± 126NDND
T1•HCl>400>400529 ± 108NDND
T2•HCl266 ± 8313 ± 40444 ± 22NDND
a SEM: Standard error of mean of at least three independent assays; b SEM: standard error of mean of two independent assays; c SI for amastigote forms; ND: not defined.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Amado, P.S.M.; Costa, I.C.C.; Paixão, J.A.; Mendes, R.F.; Cortes, S.; Cristiano, M.L.S. Synthesis, Structure and Antileishmanial Evaluation of Endoperoxide–Pyrazole Hybrids. Molecules 2022, 27, 5401. https://doi.org/10.3390/molecules27175401

AMA Style

Amado PSM, Costa ICC, Paixão JA, Mendes RF, Cortes S, Cristiano MLS. Synthesis, Structure and Antileishmanial Evaluation of Endoperoxide–Pyrazole Hybrids. Molecules. 2022; 27(17):5401. https://doi.org/10.3390/molecules27175401

Chicago/Turabian Style

Amado, Patrícia S. M., Inês C. C. Costa, José A. Paixão, Ricardo F. Mendes, Sofia Cortes, and Maria L. S. Cristiano. 2022. "Synthesis, Structure and Antileishmanial Evaluation of Endoperoxide–Pyrazole Hybrids" Molecules 27, no. 17: 5401. https://doi.org/10.3390/molecules27175401

Article Metrics

Back to TopTop