Next Article in Journal
Plant Essential Oils as Healthy Functional Ingredients of Nutraceuticals and Diet Supplements: A Review
Next Article in Special Issue
Recent Trends in the Development of Novel Metal-Based Antineoplastic Drugs
Previous Article in Journal
New Copper(II)-L-Dipeptide-Bathophenanthroline Complexes as Potential Anticancer Agents—Synthesis, Characterization and Cytotoxicity Studies—And Comparative DNA-Binding Study of Related Phen Complexes
Previous Article in Special Issue
Highlights of New Strategies to Increase the Efficacy of Transition Metal Complexes for Cancer Treatments
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis and Antiparasitic Activity of New Trithiolato-Bridged Dinuclear Ruthenium(II)-arene-carbohydrate Conjugates

1
Department of Chemistry, Biochemistry and Pharmaceutical Sciences, University of Bern, Freiestrasse 3, 3012 Bern, Switzerland
2
Institute of Parasitology Vetsuisse Faculty, University of Bern, Länggass-Strasse 122, 3012 Bern, Switzerland
3
School of Chemistry, Cardiff University, Park Place, Cardiff CF103AT, UK
*
Authors to whom correspondence should be addressed.
Submission received: 28 November 2022 / Revised: 9 January 2023 / Accepted: 11 January 2023 / Published: 16 January 2023
(This article belongs to the Special Issue Metal-Based Drugs: Past, Present and Future)

Abstract

:
Eight novel carbohydrate-tethered trithiolato dinuclear ruthenium(II)-arene complexes were synthesized using CuAAC ‘click’ (Cu(I)-catalyzed azide-alkyne cycloaddition) reactions, and there in vitro activity against transgenic T. gondii tachyzoites constitutively expressing β-galactosidase (T. gondii β-gal) and in non-infected human foreskin fibroblasts, HFF, was determined at 0.1 and 1 µM. When evaluated at 1 µM, seven diruthenium-carbohydrate conjugates strongly impaired parasite proliferation by >90%, while HFF viability was retained at 50% or more, and they were further subjected to the half-maximal inhibitory concentration (IC50) measurement on T. gondii β-gal. Results revealed that the biological activity of the hybrids was influenced both by the nature of the carbohydrate (glucose vs. galactose) appended on ruthenium complex and the type/length of the linker between the two units. 23 and 26, two galactose-based diruthenium conjugates, exhibited low IC50 values and reduced effect on HFF viability when applied at 2.5 µM (23: IC50 = 0.032 µM/HFF viability 92% and 26: IC50 = 0.153 µM/HFF viability 97%). Remarkably, compounds 23 and 26 performed significantly better than the corresponding carbohydrate non-modified diruthenium complexes, showing that this type of conjugates are a promising approach for obtaining new antiparasitic compounds with reduced toxicity.

Graphical Abstract

1. Introduction

The interest in the development of metal complexes for medicinal applications increased in the middle of the 20th century after the discovery of the anticancer properties of cisplatin [1,2]. Metal-based drugs are attractive due to their great versatility in terms of metal center, oxidation state, coordination number, in addition to the nature and geometric orientation of the ligands [3]. As the use of platinum-based drugs is limited due to shortcomings like the occurrence of chemoresistance and side effects associated to their high toxicity [4,5], this encouraged the research of compounds based on other metals as alternative to platinum anticancer therapeutics [1,6,7,8]. Parallel investigations aimed to enlarge the purpose of metal complexes with the identification of additional pharmacological properties, such as antibiotic [9,10] and antiparasitic [11,12,13,14,15,16].
Ruthenium complexes were identified amid the most promising non-platinum chemotherapeutic alternatives [17,18]. The ruthenium(II)-arene scaffold has been declined in a myriad of compounds aimed to improve anticancer activity and selectivity [19,20,21,22,23,24,25], but also targeting other therapeutic applications [11,13,16,26].
A particular class of compounds containing this unit are the trithiolato-bridged dinuclear ruthenium(II)-arene complexes (AC in Figure 1), which show not only high antiproliferative activity against cancer cells [27], but also promising antiparasitic properties [28,29]. The structure of these complexes is based on a trigonal bipyramidal Ru2S3 framework, with two ruthenium(II)-arene half-sandwich units. Two types of complexes can be distinguished, “mixed” (at least one of the bridge thiols is different, A in Figure 1) and “symmetric” (the three bridge thiols are identical, B and C in Figure 1) [27]. Former studies on Toxoplasma gondii [28], Neospora caninum [29] and Trypanosoma brucei [30] identified high antiparasitic activity for some of these diruthenium compounds. For example, compounds AC (Figure 1) inhibit T. gondii tachyzoites proliferation with IC50 values in nanomolar range (down to 1.2 nM for A).
T. gondii is an obligate intracellular protozoan parasite of the phylum Apicomplexa that causes infections of medical and veterinary significance in humans and animals [31,32]. Infection is usually asymptomatic in immunocompetent individuals, but it may cause severe complications or even be fatal in immunocompromised patients [33]. Current common treatments for toxoplasmosis are not specific, require prolonged courses and have toxic side effects, and consequently, new therapeutic solutions are needed [33,34,35,36]. Unlike other pathogens, T. gondii has adapted to replicate in all nucleated cells of a wide range of vertebrates, regardless of their cellular metabolism, and thus displays an exceptional metabolic robustness [37,38]. T. gondii is auxotrophic for several metabolites including purines, polyamines, cholesterol and choline [37]. Accordingly, tackling the parasite auxotrophies and metabolic peculiarities can constitute an interesting therapeutic strategy [37].
Carbohydrates contribute to cell-cell recognition and adhesion, have a crucial role in cellular energy supply, and can bind to specific proteins (e.g., lectins, glucose transporters, and glycoenzymes). Consequently, their conjugation to metal complexes appears as a rational choice for drug design as it can promote biocompatibility and increase water solubility. Carbohydrate-metal hybrids show promise not only in medicinal chemistry [39,40,41] but also in catalysis [42,43]. Apart the metal, its oxidation state and coordination mode [44,45,46], various structural adjustments were considered as the type of the carbohydrate [47,48,49,50,51], its substitution position [52], and the presence and nature of the protecting groups [53,54,55]. The cancer cells glucose metabolism can be exploited for targeted therapy [56], and consequently, glycoconjugates of various metal complexes were explicitly designed for selective uptake by cells overexpressing glucose transporters [57,58,59]. In this context, the potential of ruthenium complexes containing carbohydrate-functionalized ligands has been extensively studied especially on cancer cells [60,61,62,63,64,65,66,67,68,69,70,71,72,73], and some representative examples are presented in Figure 1. Apart cancer-specific treatment [39,40,41,60,74,75,76,77], alternative utilizations of metal-carbohydrate hybrids, as for example antiparasitic therapy, also received a lot of interest [78,79,80] (Figure 1).
For example Ru(II)-arene complexes as D [65,66], with a carbohydrate-derived phosphorus-containing ligand, E [67] bearing a mannose fragment as a diamino-bidentate leg ligand, F [68,69], with a galactose fragment N-coordinated via a nitrile group, and G [71], containing a glucosyl functionalized 1,2,3-triazolylidene N-heterocyclic carbene ligand, exhibited promising antiproliferative activity on various cancer cells. Complexes like H [72], with methyl mannose or glucose units attached to a pyridyl-2-triazole bidentate ligand, were shown to exploit the glucose transporters for cellular uptake in cancer cells. For Ru(II) half-sandwich complexes like I [70] and J [73], the presence and nature of the protective groups proved to be essential for the biological activity. The high affinity of the malaria parasite for glucose was targeted using the ferrocenyl-glucose conjugate K [79], with moderate antimalarial activity in vitro in both Plasmodium falciparum chloroquine-resistant and non-resistant strains. Carbohydrate-ferrocenyltriazole conjugate L [81], exhibited antibacterial activity against both Gram-positive and Gram-negative pathogens, and triazole bridged ferrocene-selenoribose conjugate M [82] was cytotoxic on cancer cells.
This study continues the quest for trithiolato-bridged dinuclear ruthenium(II)-arene compounds as potential anti-Toxoplasma compounds with improved therapeutic value (in terms of antiparasitic efficacy/host cell toxicity balance) by exploiting the conjugate strategy and the parasite auxotrophies and specific metabolic needs. The investigation of carbohydrate metabolism in T. gondii has received a lot of interest [83,84,85,86] and considering the high energetic demand accompanying parasite growth and proliferation, carbohydrates can constitute an appealing choice among the metabolites able to promote the internalization of the organometallic unit in the parasite.
The synthesis of trithiolato diruthenium complexes is generally straightforward and efficient [87,88,89], this scaffold being robust to chemical modification and easily adaptable to the conjugate strategy as demonstrated by the various series of hybrids with peptides [90], drugs [91,92], fluorophores [89,93] or metabolites [93]. Ester and amide couplings [89,94], but also CuAAC (Cu(I)-catalyzed azide-alkyne cycloaddition) click reactions [92,93] proved to be useful tools for the functionalization of the diruthenium trithiolato unit at the level of the bridge thiols. CuAAC offer the advantage of mild reaction conditions, compatible with various ligands [44,46,71,72,73,95,96,97] but also with organometallics [81,98,99,100,101,102], and enables the construction of libraries of compounds [103,104,105]. Additionally, trithiolato diruthenium(II)-arene compounds suitably substituted with alkyne or azide groups were already used in CuAAC reactions for obtaining conjugates with molecules of interest e.g., various nucleic bases or drugs [92,93].
The nature of the carbohydrate (acetyl protected glucose or galactose) and the type and length of the linker between the two units were addressed as sources of variability. The new diruthenium hybrids and intermediates were screened in vitro against T. gondii tachyzoites expressing β-galactosidase (T. gondii β-gal) grown in human foreskin fibroblasts (HFF) with complementary assessment of HFF host cells viability. Compounds with promising antiparasitic activity and selectivity were then subjected to dose-response (IC50) determination on T. gondii β-gal and toxicity assessment on HFF at 2.5 M concentration.

2. Results and Discussions

2.1. Synthesis

2.1.1. Synthesis of the Dinuclear Ruthenium(II)-arene Intermediates 29

Alkyne and azide partners are needed for the CuAAC reactions, and when appropriately substituted, both the diruthenium moiety and the carbohydrate can play either role. With this aim, various diruthenium and carbohydrate intermediates were synthesized.
The dithiolato derivative 1 [106] (obtained from the ruthenium dimer ([(η6-p-MeC6H4Pri)RuCl]2Cl2) and 4-tert-butylbenzenemethanethiol) was reacted with a second thiol (4-mercaptophenol, 4-aminobenzenthiol, 2-(4-mercaptophenyl)acetic acid, and 2-mercaptobenzyl alcohol, respectively) to provide the trithiolato-bridged dinuclear ruthenium compounds 25, as previously reported (Scheme 1) [87,88,89].
Intermediates 24 can be modified using ester and amide coupling reactions as previously described [89,92,93,94]. The alkyne ester 6 was obtained in moderate yield (47%) by reacting 2 with 5-hexynoic acid using EDCI (N-(3-dimethylaminopropyl)-N′-ethylcarbodiimide hydrochloride) as coupling agent, in basic conditions (DMAP, 4-(dimethylamino)-pyridine) (Scheme 2, top). 5-Hexynoic acid was also reacted with the amino diruthenium derivative 3 using EDCI and HOBt (1-hydroxybenzotriazole) as coupling agents, in basic conditions (DIPEA, N,N-diisopropylethylamine), to afford the amido alkyne compound 7 as reported [93] (Scheme 2, top). Similar reaction conditions were used for the synthesis of amide 8 from carboxylic acid diruthenium derivative 4 and propargylic amine as formerly described (Scheme 2, bottom) [92,93].
The azide trithiolato diruthenium derivative 9 (Scheme 3), was obtained following a two steps pathway starting from alcohol 5 using a reported protocol [93]. First, the hydroxy group was activated by mesylation (MsCl, methanesulfonyl chloride) in basic conditions (TEA, triethylamine), followed by the nucleophilic substitution with azide (NaN3).

2.1.2. Synthesis of the Azide and Alkyne Functionalized Carbohydrate Intermediates 1018

Appropriate carbohydrate derivatives bearing azide and alkyne groups were also synthesized (Scheme 4 and Scheme 5). Azido glucose compound 10 (2,3,4,6-tetra-O-acetyl-β-D-glucopyranosyl azide), was synthesized from commercially available 2,3,4,6-tetra-O-acetyl-β-D-glucopyranosyl bromide following a literature protocol (Scheme 4) [107]. The reaction was realized with TMS-N3 (trimethylsilylazide) in THF in the presence of TBAF (tetrabutylammonium fluoride) in catalytic amounts, and 10 was isolated in moderate yield (51%).
The azide compounds 1416 were obtained following a two-step procedure previously described [52,55,108] (Scheme 5, top). First, β-D-glucose pentaacetate and β-D-galactose pentaacetate were glycosylated with 2-bromoethanol and 4-bromo-1-butanol. The reactions were realized in the presence of BF3∙Et2O (boron trifluoride diethyl etherate) as Lewis acid catalyst [52] and afforded the ether glycosides 1113 in low to moderate yields (32, 41, and 47%, respectively). In the second step the bromine atom on the pending chain of 1113 was substituted with azide (NaN3) [55,108], derivatives 1416 being isolated in 47, 74% and quantitative yields, respectively.
The alkyne functionalized carbohydrates 17 and 18 were synthesized (Scheme 5, bottom) from β-D-glucose pentaacetate and, respectively, from β-D-galactose pentaacetate and 4-pentyn-1-ol in the presence of BF3∙Et2O [109] and were isolated in medium yields (64 and 46%).

2.1.3. Synthesis of the Carbohydrate Functionalized Trithiolato-Bridged Dinuclear Ruthenium(II)-arene Complexes 1926

The carbohydrate units were attached to the trithiolato diruthenium scaffold via click 1,3-dipolar cycloadditions using adapted protocols [110,111,112], in the presence of CuSO4 as catalyst and sodium ascorbate as a reducing agent, in DMF under inert conditions. Complexes 69, bearing either alkyne or azide pendant group, were reacted with the appropriately functionalized carbohydrate derivatives 10 and 1418 (Scheme 6, Scheme 7, Scheme 8 and Scheme 9) affording eight new trithiolato diruthenium conjugates 1926.
Thus, alkyne functionalized diruthenium compounds 6 and 7 were reacted with glucose derivative 10 presenting an azide group directly anchored to the glucopyranosyl ring (Scheme 6). Amide conjugate 20 was isolated in good yield (72%), while difficulties were encountered in the purification of ester analogue 19 which was recovered in poorer yield (28%).
Other glycoconjugates were synthesized using alkyne intermediate 7 (Scheme 7) and two types of modifications were envisioned: (i) the nature of the carbohydrate (glucose in 21 vs. galactose in 22), and (ii) the presence of spacers of different length between the azide group and the glucopyranosyl ring (galactose derivatives 22 and 23). Conjugates 2123 were isolated in good yields of 66, 65 and 74%, respectively. Neither the steric hindrance nor the nature of the carbon atom on which the azide group was anchored (10 vs. 14) play a key role on the yield (20 vs. 2123). Similarly, the reaction of the diruthenium propargyl amide derivative 8 with galactose azide 14 afforded conjugate 24 in 75% yield (Scheme 8).
The trithiolato dinuclear intermediate 9, presenting an azide in benzylic position on one of the bridge thiols, was reacted with glucose and galactose alkyne derivatives 17 and 18 (Scheme 9), affording carbohydrate conjugates 25 and 26 isolated in moderate yields of 51 and 63%, respectively.
All compounds were fully characterized by 1H, 13C nuclear magnetic resonance (NMR) spectroscopy, high resolution electrospray ionization mass spectrometry (HR ESI-MS) and elemental analysis (see Supplementary Materials). The obtainment of the triazole connector between the diruthenium unit and the carbohydrates was undoubtedly demonstrated by the 1H and 13C NMR spectra of the conjugates 19–26 by the signals corresponding to the proton of the triazole cycle at 7.74–8.66 ppm and of the corresponding carbon at 120.3–124.6 ppm. The absence of the signals corresponding to the proton of the monosubstituted alkyne (at 1.93–2.49 in compounds 6–8, 17 and 18) in the 1H NMR spectra of conjugates 19–26 further confirm the obtainment of the hybrid molecules. Mass spectrometry corroborated the spectroscopic data with the trithiolato diruthenium glucose and galactose conjugates 1926 showing molecular ion peaks corresponding to [M-Cl]+ ions.

2.1.4. Stability of the Compounds

For the assessment of the biological activity, the compounds were prepared as stock solutions in dimethylsulfoxide (DMSO). Similar to former reports [88,89,94], the 1H NMR spectra of the functionalized diruthenium complexes 6, 7, 20, 22, 25 and 26 in DMSO-d6, recorded at 25 °C 5 min and more than 1 month after sample preparation showed no significant modifications (see Figure S1 in the Supplementary Materials), demonstrating a very good stability of the compounds in this highly complexing solvent.
Compound 19 has an ester linker that can potentially be hydrolyzed in cell growth media. Comparable conjugates with fluorophores (coumarin and BODIPY) linked through ester bonds to the trithiolato diruthenium unit were recently studied [89,93]. Only very limited solvolysis of the ester bonds was noticed after 168 h for some compounds, and it was concluded that the fluorophore diruthenium conjugates exhibit high stability in the conditions used for the biological evaluations. Therefore, it was assumed that compound 19 is appropriately stable for the first in vitro biological activity evaluation.

2.2. Assessment of the In Vitro Activity against T. Gondii β-gal and Human Foreskin Fibroblast Host Cells

2.2.1. Primary Screening

The biological activity of the carbohydrate azides and alkyne derivatives 1416 and, respectively, 17 and 18 was not measured as these compounds were not isolated pure. Glucose and galactose conjugates 1926, glucose azide derivative 10 and diruthenium alkyne intermediate 6 were assessed for their in vitro biological activity in inhibiting proliferation of T. gondii β-gal, a transgenic strain that constitutively expresses β-galactosidase, and for toxicity to HFF (human foreskin fibroblast) used as host cells. The compounds were applied to infected or non-infected HFF cultures for 72 h and at concentrations of 0.1 and 1 µM, the results being summarized in Table 1 and Figure 2. The viability of treated HFF was measured by the alamarBlue metabolic assay, and the proliferation of T. gondii β-gal was quantified by the β-galactosidase colorimetric test. In both cases, results are expressed as percentage (%) compared to control parasitic and host cells treated with 0.1% DMSO for which proliferation and viability were set to 100% (Table 1).
The trithiolato diruthenium complexes 25 and 9, and alkyne intermediates 7 and 8 were evaluated previously against T. gondii β-gal under similar conditions [88,89,92,93], and the corresponding values are shown in Table 1 and Figure 2 for comparison. The new alkyne ester derivative 6 impacted T. gondii proliferation but affected significantly less the viability of HFF compared to its diruthenium hydroxy intermediates 2. The glucose azide derivative 10 exhibited neither antiparasitic activity nor host cell toxicity at both tested concentrations.
In the first screening, the eight carbohydrate conjugates 1926, applied at 1 µM, did not impair host cell viability. Apart from glucose conjugate 25, all the dyads nearly abolished parasite proliferation when applied at 1 µM. Though, apart from glucose conjugate 19, all hybrid molecules had only a limited effect on T. gondii β-gal at 0.1 µM.
Both the type of carbohydrate and the nature and length of the linker influenced the antiparasitic efficacy and cytotoxicity of the conjugates, but no clear straightforward trends could be identified regarding the relationship between the structural elements and biological activity. For example, when the conjugates are applied at 0.1 µM, some differences in anti-Toxoplasma efficacy are observed. For instance, glucose ester derivative 19 is significantly more active on T. gondii compared to the amide analogue 20. Galactose functionalized compound 22 is more efficient in inhibiting the parasite proliferation compared to the corresponding glucose derivative 21, while for the same carbohydrate an increase of the linker length has a negative effect on the antiparasitic activity (galactose conjugates 22 and 23).

2.2.2. IC50 Values against T. gondii β-gal Tachyzoites and HFF Toxicity at 2.5 µM

For a compound to be selected for the second screening, two criteria had to be met simultaneously: (i) when the compound was applied at 1 µM, T. gondii β-gal growth was inhibited by 90% or more compared to control treated with 0.1% DMSO only, and (ii) HFF host cell viability was not impaired by more than 50% for a compound applied at 1 µM. Based on the results of the primary screening, glucose and galactose dyads 1924 and 26 were selected. Pyrimethamine, currently used for the treatment of toxoplasmosis, and which inhibited the proliferation of T. gondii β-gal tachyzoites with an IC50 value of 0.326 µM and did not affect HFF viability at 2.5 µM (Table 2), was used as reference compound. The selection also included the diruthenium intermediate compounds 2, 3 and 5 with free OH or NH2 groups, along with two diruthenium alkyne ester and amide compounds 6 and 7, and diruthenium azide 9. The results are summarized in Table 2.
The IC50 values and the cytotoxicity of the diruthenium compounds 2, 3, 5, 7 and 9 were measured previously [88,89,92,93]. For these diruthenium intermediates the IC50 values ranged from 0.025 µM (6) to 0.153 µM (3). However, all intermediates also strongly affected the viability of HFF when applied at 2.5 µM, the most cytotoxic being compounds 5, 6 and 9.
Glucose conjugates 19 and 21 exhibited low IC50 values (0.018 and 0.087 µM, respectively), but were toxic to host cells at 2.5 µM (HFF viability was reduced to 29% for 19 and abolished for 21). Glucose hybrid 20 exhibited antiparasitic activity (IC50 = 0.110 µM) but also medium cytotoxicity (HFF viability of 77%). Galactose and glucose dyads 22 and 24 had only modest antiparasitic activity (IC50 values of 0.294 and 0.328 µM, respectively, comparable with those obtained for pyrimethamine) while being moderately toxic to HFF at 2.5 µM (73 and 66%, significantly more cytotoxic compared to the standard pyrimethamine).
Galactose conjugates 23 and 26 were the most promising of the series exhibiting not only high efficacy in inhibiting T. gondii β-gal proliferation (IC50 values of 0.032 and 0.153 µM, 10-fold and 2-fold lower compared to pyrimethamine, IC50 = 0.326 µM), but also low cytotoxicity on the host cells when applied at 2.5 µM (HFF viability 92 and 97%, respectively).
Interestingly, both glucose and galactose hybrids 20 and 23 affected the HFF viability less than the diruthenium alkyne intermediate 7 from which they were obtained by click reactions. A similar result was also obtained for the galactose conjugate 26 compared to the diruthenium azide parent 9.
The number of conjugates considered in this study is too limited to allow proper SAR observations. Nevertheless, apart from the conjugation with protected carbohydrates, other structural features of the dyads (as the nature and length of the linker between the two units), appear to strongly influence the biological activity, and a fine structural tuning is needed to obtain compounds with good pharmacological properties in terms of safety/anti-toxoplasma efficacy balance.
Further studies are necessary for the identification of the mode of action of trithiolato diruthenium compounds. For some other types of dinuclear Ru(II)-arene complexes reported in the literature, interactions with DNA and oligonucleotide sequences were identified [113,114,115,116,117,118]. However, unlike other Ru(II)-arene complexes presenting labile chlorine, carboxylate or monodentate N-coordinated ligands, the trithiolato diruthenium complexes do not hydrolyze and are stable in the presence of most biomolecules such as amino acids and DNA [27]. Furthermore, a recent study revealed only weak interactions via H-bonding nucleobase-pairing between trithiolato diruthenium nucleobase conjugates and the respective complementary nucleic bases [93]. In the presence of some trithiolato diruthenium complexes the oxidation of cysteine (Cys) and glutathione (GSH) to form cystine and GSSG, respectively, was observed [119,120]. TEM (transmission electron microscopy) studies of different protozoan parasites (Toxoplasma gondii, Neospora caninum, Trypanosoma brucei) treated with trithiolato dinuclear ruthenium(II)-arene complexes revealed alterations in the mitochondrial ultrastructure indicating this parasite organelle as potential target [29]. Noteworthy, trithiolato diruthenium conjugates with coumarin and BODIPY fluorophores [89,121] induced analogous outcome on parasite mitochondrion.

3. Materials and Methods

3.1. Chemistry

The chemistry experimental part, with full description of synthetic procedures and characterization data for all compounds are presented in the Supplementary Materials.

3.2. Biological Evaluation

3.2.1. Cell and Parasite Culture

All tissue culture media were purchased from Gibco-BRL, and biochemical agents from Sigma-Aldrich. Human foreskin fibroblasts (HFF) were obtained from the American Type Culture Collection (ATCC) and maintained in complete culture medium consisting in DMEM (Dulbecco’s Modified Eagle’s Medium) supplemented with 10% fetal calf serum (FCS, Gibco-BRL, Waltham, MA, USA) and antibiotics as previously described [122]. Transgenic T. gondii β-gal tachyzoites (expressing the β-galactosidase gene from Escherichia coli) from RH strain were kindly provided by Prof. David Sibley (Washington University, St. Louis, MO, USA) and were maintained by passages in HFF cultures as previously described [122,123].

3.2.2. In Vitro Activity Assessment against T. Gondii Tachyzoites and Human Foreskin Fibroblasts

The screening sequence for the compounds was described in previous reports [88]. All compounds were prepared as 1 mM stock solutions from powder in dimethyl sulfoxide (DMSO, Sigma, St. Louis, MO, USA). For in vitro activity and cytotoxicity assays, HFF were seeded at 5 × 103/well in 96 well plates and allowed to grow to confluence in complete culture medium at 37 °C and 5% CO2. Transgenic T. gondii β-gal tachyzoites were freshly isolated from infected cultures as described [122], and 96-well plates containing HFF monolayer were infected with 1 × 103 tachyzoites/well.
In a primary screening, each compound was evaluated at two concentrations 0.1 and 1 µM and added to the media prior to the infection as previously described [94]. Control non-infected non-treated HFF cultures and T. gondii β-gal infected but not-treated cultures were cultivated in complete medium containing 0.01 or 0.1% DMSO. The 96-well plates were incubated for 72 h at 37 °C/5% CO2 as previously described [94].
For the IC50 determination on T. gondii β-gal, eight serial concentrations ranging from 7 nM to 1 μM were tested for each selected compound as previously described [89,92,93] and the β-galactosidase assay was performed as reported [122]. Briefly, infected HFF cultures in 96-well plates were lysed with PBS containing 0.05% Triton X-100. Then the substrate chlorophenolred-β-D-galactopyranoside (CPRG; Roche Diagnostics, Rotkreuz, Switzerland) was added in a final concentration of 0.5 mM. Absorption was measured at 570 nm wavelength using an EnSpire® multimode plate reader (PerkinElmer, Inc., Waltham, MA, USA).
All calculations were performed using the corresponding software tool contained in the Excel software package (Microsoft, Redmond, WA, USA). Cytotoxicity assays using uninfected confluent HFF host cells were performed by the alamarBlue assay as previously reported [124]. Confluent HFF monolayers in 96-well plates were exposed to 0.1, 1 and 2.5 μM of each compound and incubated for 72 h at 37 °C/5% CO2. Then the medium was removed, the plates were washed once with PBS and 200 μL of resazurin (1:200 dilution in PBS) were added to each well. Plates were measured at excitation wavelength 530 nm and emission wavelength 590 nM using an EnSpire® multimode plate reader (PerkinElmer, Inc.). Fluorescence was measured at two different time points: T0 as starting timepoint and T5h as at 5 h later. Relative fluorescence units were calculated from time points with linear increases.

4. Conclusions

This study was focused on the synthesis and in vitro anti-Toxoplasma activity evaluation of eight new trithiolato-bridged arene-ruthenium(II) carbohydrate conjugates. Acetyl protected glucose and galactose moieties were pended on the diruthenium unit on one of the bridging thiols using CuAAC click reactions and connectors of several types and lengths to obtain the carbohydrate dyads. In the first screening, none of the conjugates affected the validity of host cells at 1 µM, suggesting reduced toxicity, and seven carbohydrate-diruthenium hybrids applied at 1 µM inhibited T. gondii β-gal growth by more than 90%. The second screening (IC50 values and toxicity to HFF after exposure to 2.5 µM) led to the identification of two promising acetyl protected galactose functionalized compounds 23 and 26. Both conjugates not only exceeded (up to 10-fold) the anti-Toxoplasma efficacy of the standard drug pyrimethamine for similar level of toxicity to HFF, but also exhibited a significantly better antiparasitic activity/cytotoxicity balance compared to the corresponding carbohydrate non-modified diruthenium complexes.
The type and length of the linker between the diruthenium core and the carbohydrate unit significantly influenced the biological activity, and fine structural adjustments could further increase the anti-Toxoplasma efficacy of this type of carbohydrate conjugates. In addition, the nature of the carbohydrate and the presence/absence of protecting groups is known to strongly affect the biological activity of conjugates carbohydrate-organometallic complex [53,70,73]. Thus, the use of other carbohydrates bearing, or no protective groups is also considered.
This study showed that carbohydrate conjugation to trithiolato-diruthenium complexes is a promising strategy for obtaining novel organometallic compounds with high antiparasitic efficacy and reduced host cell cytotoxicity.

Supplementary Materials

The following supporting information can be downloaded at https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/molecules28020902/s1. Synthetical procedures; Synthesis of the trithiolato-bridged dinuclear ruthenium(II)-arene intermediates 29; Synthesis of the azide and alkyne functionalized carbohydrate intermediates 1018; Synthesis of the carbohydrate functionalized trithiolato-bridged dinuclear ruthenium(II)-arene complexes 1926. Figure S1. 1H NMR Spectra of 6, 7, 20, 22, 25 and 26 recorded in DMSO-d6 at 25°C as function of time. References [125,126] are cited in the Supplementary Materials.

Author Contributions

Conceptualization, A.H., J.F. and E.P.; Methodology, G.B., A.H., J.F. and E.P.; Software, I.H., O.D., N.A., S.K.J., G.B. and E.P.; Validation, O.D., G.B., A.H., J.F. and E.P.; Formal Analysis, I.H., O.D., N.A., S.K.J., G.B. and E.P.; Investigation, I.H., O.D., N.A., S.K.J., G.B. and E.P.; Resources, A.H. and J.F.; Data Curation, O.D., G.B., N.A. and E.P.; Writing—Original Draft Preparation, O.D., G.B., N.A. and E.P.; Writing—Review and Editing, O.D., G.B., N.A., A.H., J.F. and E.P.; Supervision, G.B., A.H., J.F. and E.P.; Project Administration, A.H. and J.F.; Funding Acquisition, A.H. and J.F. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the Swiss Science National Foundation (SNF, Sinergia project CRSII5-173718 and project no. 310030_184662).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data is included in the article and the Supplementary Materials.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Some compounds are available from the authors upon request.

References

  1. Muhammad, N.; Guo, Z. Metal-based anticancer chemotherapeutic agents. Curr. Opin. Chem. Biol. 2014, 19, 144–153. [Google Scholar] [CrossRef] [PubMed]
  2. Farrell, N. Metal complexes as drugs and chemotherapeutic agents. In Comprehensive Coordination Chemistry II; Elsevier: Amsterdam, The Netherlands, 2004; Volume 9, pp. 809–840. [Google Scholar]
  3. Ndagi, U.; Mhlongo, N.; Soliman, M.E. Metal complexes in cancer therapy–an update from drug design perspective. Drug Des. Devel. Ther. 2017, 11, 599–616. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Makovec, T. Cisplatin and beyond: Molecular mechanisms of action and drug resistance development in cancer chemotherapy. Radiol. Oncol. 2019, 53, 148–158. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Galluzzi, L.; Senovilla, L.; Vitale, I.; Michels, J.; Martins, I.; Kepp, O.; Castedo, M.; Kroemer, G. Molecular mechanisms of cisplatin resistance. Oncogene 2012, 31, 1869–1883. [Google Scholar] [CrossRef] [Green Version]
  6. Karges, J.; Stokes, R.W.; Cohen, S.M. Metal complexes for therapeutic applications. Trends Chem. 2021, 3, 523–534. [Google Scholar] [CrossRef]
  7. Ong, Y.C.; Gasser, G. Organometallic compounds in drug discovery: Past, present and future. Drug Discov. Today Technol. 2020, 37, 117–124. [Google Scholar] [CrossRef]
  8. de Oliveira Silva, D. Ruthenium compounds targeting cancer therapy. In Frontiers in Anti-Cancer Drug Discovery; Atta-ur-Rahman, I., Choudhary, M., Eds.; Bentham Science: Sharjah, United Arab Emirates, 2014; Volume 4, pp. 88–156. [Google Scholar]
  9. Frei, A.; Zuegg, J.; Elliott, A.G.; Baker, M.; Braese, S.; Brown, C.; Chen, F.; Dowson, C.G.; Du, G. Metal complexes, an untapped source of antibiotic potential? Antibiotics 2020, 9, 90. [Google Scholar] [CrossRef] [Green Version]
  10. Frei, A.; Zuegg, J.; Elliott, A.G.; Baker, M.; Braese, S.; Brown, C.; Chen, F.; Dowson, C.G.; Dujardin, G.; Jung, N.; et al. Metal complexes as a promising source for new antibiotics. Chem. Sci. 2020, 11, 2627–2639. [Google Scholar] [CrossRef] [Green Version]
  11. Gambino, D.; Otero, L. Design of prospective antiparasitic metal-based compounds including selected organometallic cores. Inorg. Chim. Acta 2018, 472, 58–75. [Google Scholar] [CrossRef]
  12. Gambino, D.; Otero, Á.L. Metal compounds in the development of antiparasitic agents: Rational design from basic chemistry to the clinic. Met. Ions Life Sci. 2019, 19, 331–358. [Google Scholar] [CrossRef]
  13. Mbaba, M.; Golding, T.M.; Smith, G.S. Recent advances in the biological investigation of organometallic platinum-group metal (Ir, Ru, Rh, Os, Pd, Pt) complexes as antimalarial agents. Molecules 2020, 25, 5276. [Google Scholar] [CrossRef] [PubMed]
  14. Sánchez-Delgado, R.A.; Anzellotti, A. Metal complexes as chemotherapeutic agents against tropical diseases: Trypanosomiasis, malaria and leishmaniasis. Mini Rev. Med. Chem. 2004, 4, 23–30. [Google Scholar] [CrossRef] [PubMed]
  15. Navarro, M.; Gabbiani, C.; Messori, L.; Gambino, D. Metal-based drugs for malaria, trypanosomiasis and leishmaniasis: Recent achievements and perspectives. Drug Discov. Today 2010, 15, 1070–1078. [Google Scholar] [CrossRef] [PubMed]
  16. Ong, Y.C.; Roy, S.; Andrews, P.C.; Gasser, G. Metal compounds against neglected tropical diseases. Chem. Rev. 2019, 119, 730–796. [Google Scholar] [CrossRef] [PubMed]
  17. Bergamo, A.; Gaiddon, C.; Schellens, J.H.M.; Beijnen, J.H.; Sava, G. Approaching tumour therapy beyond platinum drugs: Status of the art and perspectives of ruthenium drug candidates. J. Inorg. Biochem. 2012, 106, 90–99. [Google Scholar] [CrossRef] [PubMed]
  18. Alessio, E.; Messori, L. NAMI-A and KP1019/1339, two iconic ruthenium anticancer drug candidates face-to-face: A case story in medicinal inorganic chemistry. Molecules 2019, 24, 1995. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Weiss, A.; Berndsen, R.H.; Dubois, M.; Müller, C.; Schibli, R.; Griffioen, A.W.; Dyson, P.J.; Nowak-Sliwinska, P. In vivo anti-tumor activity of the organometallic ruthenium(II)-arene complex [Ru(η6-p-cymene)Cl2(pta)] (RAPTA-C) in human ovarian and colorectal carcinomas. Chem. Sci. 2014, 5, 4742–4748. [Google Scholar] [CrossRef] [Green Version]
  20. Aird, R.E.; Cummings, J.; Ritchie, A.A.; Muir, M.; Morris, R.E.; Chen, H.; Sadler, P.J.; Jodrell, D.I. In vitro and in vivo activity and cross resistance profiles of novel ruthenium(II) organometallic arene complexes in human ovarian cancer. Br. J. Cancer 2002, 86, 1652–1657. [Google Scholar] [CrossRef] [Green Version]
  21. Su, W.; Tang, Z.; Li, P. Development of arene ruthenium antitumor complexes. Mini Rev. Med. Chem. 2016, 16, 787–795. [Google Scholar] [CrossRef]
  22. Su, W.; Li, Y.; Li, P. Design of Ru-arene complexes for antitumor drugs. Mini Rev. Med. Chem. 2018, 18, 184–193. [Google Scholar] [CrossRef]
  23. Golbaghi, G.; Castonguay, A. Rationally designed ruthenium complexes for breast cancer therapy. Molecules 2020, 25, 265. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Biersack, B. Anticancer activity and modes of action of (arene) ruthenium(II) complexes coordinated to C-, N-, and O-ligands. Mini Rev. Med. Chem. 2016, 16, 804–814. [Google Scholar] [CrossRef] [PubMed]
  25. Zheng, K.; Wu, Q.; Ding, Y.; Mei, W. Arene ruthenium(II) complexes: The promising chemotherapeutic agent in inhibiting the proliferation, migration and invasion. Mini-Rev. Med. Chem. 2016, 16, 796–803. [Google Scholar] [CrossRef] [PubMed]
  26. Laurent, Q.; Batchelor, L.K.; Dyson, P.J. Applying a Trojan horse strategy to ruthenium complexes in the pursuit of novel antibacterial agents. Organometallics 2018, 37, 915–923. [Google Scholar] [CrossRef]
  27. Furrer, J.; Süss-Fink, G. Thiolato-bridged dinuclear arene ruthenium complexes and their potential as anticancer drugs. Coord. Chem. Rev. 2016, 309, 36–50. [Google Scholar] [CrossRef]
  28. Basto, A.P.; Müller, J.; Rubbiani, R.; Stibal, D.; Giannini, F.; Süss-Fink, G.; Balmer, V.; Hermphill, A.; Gasser, G.; Furrer, J. Characterization of the activities of dinuclear thiolato-bridged arene ruthenium complexes against Toxoplasma gondii. Antimicrob. Agents Chemother. 2017, 61, e01031-17. [Google Scholar] [CrossRef] [Green Version]
  29. Basto, A.P.; Anghel, N.; Rubbiani, R.; Müller, J.; Stibal, D.; Giannini, F.; Süss-Fink, G.; Balmer, V.; Gasser, G.; Furrer, J.; et al. Targeting of the mitochondrion by dinuclear thiolato-bridged arene ruthenium complexes in cancer cells and in the apicomplexan parasite Neospora caninum. Metallomics 2019, 11, 462–474. [Google Scholar] [CrossRef] [Green Version]
  30. Jelk, J.; Balmer, V.; Stibal, D.; Giannini, F.; Süss-Fink, G.; Bütikofer, P.; Furrer, J.; Hemphill, A. Anti-parasitic dinuclear thiolato-bridged arene ruthenium complexes alter the mitochondrial ultrastructure and membrane potential in Trypanosoma brucei bloodstream forms. Exp. Parasitol. 2019, 205, 107753. [Google Scholar] [CrossRef]
  31. Hill, D.; Dubey, J.P. Toxoplasma gondii: Transmission, diagnosis and prevention. Clin. Microbiol. Infect. 2002, 8, 634–640. [Google Scholar] [CrossRef] [Green Version]
  32. Tenter, A.M.; Heckeroth, A.R.; Weiss, L.M. Toxoplasma gondii: From animals to humans. Int. J. Parasitol. 2000, 30, 1217–1258. [Google Scholar] [CrossRef]
  33. Konstantinovic, N.; Guegan, H.; Stäjner, T.; Belaz, S.; Robert-Gangneux, F. Treatment of toxoplasmosis: Current options and future perspectives. Food Waterborne Parasitol. 2019, 15, e00036. [Google Scholar] [CrossRef]
  34. Alday, P.H.; Doggett, J.S. Drugs in development for toxoplasmosis: Advances, challenges, and current status. Drug Des. Dev. Ther. 2017, 11, 273–293. [Google Scholar] [CrossRef] [Green Version]
  35. Deng, Y.; Wu, T.; Zhai, S.Q.; Li, C.H. Recent progress on anti-Toxoplasma drugs discovery: Design, synthesis and screening. Eur. J. Med. Chem. 2019, 183, 111711. [Google Scholar] [CrossRef] [PubMed]
  36. Antczak, M.; Dzitko, K.; Dlugonska, H. Human toxoplasmosis-Searching for novel chemotherapeutics. Biomed. Pharmacother. 2016, 82, 677–684. [Google Scholar] [CrossRef]
  37. Coppens, I. Exploitation of auxotrophies and metabolic defects in Toxoplasma as therapeutic approaches. Int. J. Parasitol. 2014, 44, 109–120. [Google Scholar] [CrossRef] [PubMed]
  38. Blume, M.; Seeber, F. Metabolic interactions between Toxoplasma gondii and its host. F1000Research 2018, 7, 1719. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Kenny, R.G.; Marmion, C.J. Toward multi-targeted platinum and ruthenium drugs—A new paradigm in cancer drug treatment regimens? Chem. Rev. 2019, 119, 1058–1137. [Google Scholar] [CrossRef]
  40. Bononi, G.; Iacopini, D.; Cicio, G.; Di Pietro, S.; Granchi, C.; Di Bussolo, V.; Minutolo, F. Glycoconjugated metal complexes as cancer diagnostic and therapeutic agents. ChemMedChem 2021, 16, 30–64. [Google Scholar] [CrossRef]
  41. Pettenuzzo, A.; Pigot, R.; Ronconi, L. Metal-based glycoconjugates and their potential in targeted anticancer chemotherapy. Metallodrugs 2015, 1, 36–61. [Google Scholar] [CrossRef] [Green Version]
  42. Byrne, J.P.; Musembi, P.; Albrecht, M. Carbohydrate-functionalized N-heterocyclic carbene Ru(II) complexes: Synthesis, characterization and catalytic transfer hydrogenation activity. Dalton Trans. 2019, 48, 11838–11847. [Google Scholar] [CrossRef]
  43. Pretorius, R.; Olguín, J.; Albrecht, M. Carbohydrate-functionalized 1,2,3-triazolylidene complexes for application in base-free alcohol and amine oxidation. Inorg. Chem. 2017, 56, 12410–12420. [Google Scholar] [CrossRef] [PubMed]
  44. Cucciolito, M.E.; D’Amora, A.; De Feo, G.; Ferraro, G.; Giorgio, A.; Petruk, G.; Monti, D.M.; Merlino, A.; Ruffo, F. Five-coordinate platinum(II) compounds containing sugar ligands: Synthesis, characterization, cytotoxic activity, and interaction with biological macromolecules. Inorg. Chem. 2018, 57, 3133–3143. [Google Scholar] [CrossRef] [PubMed]
  45. Annunziata, A.; Cucciolito, M.E.; Esposito, R.; Ferraro, G.; Monti, D.M.; Merlino, A.; Ruffo, F. Five-coordinate platinum(II) compounds as potential anticancer agents. Eur. J. Inorg. Chem. 2020, 2020, 918–929. [Google Scholar] [CrossRef]
  46. Annunziata, A.; Liberti, D.; Bedini, E.; Cucciolito, M.E.; Loreto, D.; Monti, D.M.; Merlino, A.; Ruffo, F. Square-planar vs. trigonal bipyramidal geometry in Pt(II) complexes containing triazole-based glucose ligands as potential anticancer agents. Int. J. Mol. Sci. 2021, 22, 8704. [Google Scholar] [CrossRef]
  47. Patra, M.; Johnstone, T.C.; Suntharalingam, K.; Lippard, S.J. A potent glucose-platinum conjugate exploits glucose transporters and preferentially accumulates in cancer cells. Angew. Chem. Int. Ed. Engl. 2016, 55, 2550–2554. [Google Scholar] [CrossRef] [Green Version]
  48. Gao, X.; Liu, S.; Shi, Y.; Huang, Z.; Mi, Y.; Mi, Q.; Yang, J.; Gao, Q. Mechanistic and biological characteristics of different sugar conjugated 2-methyl malonatoplatinum(II) complexes as new tumor targeting agents. Eur. J. Med. Chem. 2017, 125, 372–384. [Google Scholar] [CrossRef]
  49. Wu, M.; Li, H.; Liu, R.; Gao, X.; Zhang, M.; Liu, P.; Fu, Z.; Yang, J.; Zhang-Negrerie, D.; Gao, Q. Galactose conjugated platinum(II) complex targeting the Warburg effect for treatment of non-small cell lung cancer and colon cancer. Eur. J. Med. Chem. 2016, 110, 32–42. [Google Scholar] [CrossRef]
  50. Quan, L.; Lin, Z.; Lin, Y.; Wei, Y.; Lei, L.; Li, Y.; Tan, G.; Xiao, M.; Wu, T. Glucose-modification of cisplatin to facilitate cellular uptake, mitigate toxicity to normal cells, and improve anti-cancer effect in cancer cells. J. Mol. Struct. 2020, 1203, 127361. [Google Scholar] [CrossRef]
  51. Wang, H.; Yang, X.; Zhao, C.; Wang, P.G.; Wang, X. Glucose-conjugated platinum(IV) complexes as tumor-targeting agents: Design, synthesis and biological evaluation. Bioorg. Med. Chem. 2019, 27, 1639–1645. [Google Scholar] [CrossRef]
  52. Patra, M.; Awuah, S.G.; Lippard, S.J. Chemical approach to positional isomers of glucose–platinum conjugates reveals specific cancer targeting through glucose-transporter-mediated uptake in vitro and in vivo. J. Am. Chem. Soc. 2016, 138, 12541–12551. [Google Scholar] [CrossRef]
  53. Ma, J.; Liu, H.; Xi, Z.; Hou, J.; Li, Y.; Niu, J.; Liu, T.; Bi, S.; Wang, X.; Wang, C.; et al. Protected and de-protected platinum(IV) glycoconjugates with GLUT1 and OCT2-mediated selective cancer targeting: Demonstrated enhanced transporter-mediated cytotoxic properties in vitro and in vivo. Front. Chem. 2018, 6, 386. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Ma, J.; Wang, Q.; Yang, X.; Hao, W.; Huang, Z.; Zhang, J.; Wang, X.; Wang, P.G. Glycosylated platinum(IV) prodrugs demonstrated significant therapeutic efficacy in cancer cells and minimized side-effects. Dalton Trans. 2016, 45, 11830–11838. [Google Scholar] [CrossRef] [PubMed]
  55. Ma, J.; Yang, X.; Hao, W.; Huang, Z.; Wang, X.; Wang, P.G. Mono-functionalized glycosylated platinum(IV) complexes possessed both pH and redox dual-responsive properties: Exhibited enhanced safety and preferentially accumulated in cancer cells in vitro and in vivo. Eur. J. Med. Chem. 2017, 128, 45–55. [Google Scholar] [CrossRef] [PubMed]
  56. Calvaresi, E.C.; Hergenrother, P.J. Glucose conjugation for the specific targeting and treatment of cancer. Chem. Sci. 2013, 4, 2319–2333. [Google Scholar] [CrossRef] [Green Version]
  57. Medina, R.A.; Owen, G.I. Glucose transporters: Expression, regulation and cancer. Biol. Res. 2002, 35, 9–26. [Google Scholar] [CrossRef]
  58. Macheda, M.L.; Rogers, S.; Best, J.D. Molecular and cellular regulation of glucose transporter (GLUT) proteins in cancer. J. Cell. Physiol. 2005, 202, 654–662. [Google Scholar] [CrossRef] [PubMed]
  59. Szablewski, L. Expression of glucose transporters in cancers. Biochim. Biophys. Acta 2013, 1835, 164–169. [Google Scholar] [CrossRef]
  60. Fernandes, A.C. Synthesis, biological activity and medicinal applications of ruthenium complexes containing carbohydrate ligands. Curr. Med. Chem. 2019, 26, 6412–6437. [Google Scholar] [CrossRef]
  61. Iacopini, D.; Biancalana, L.; Di Pietro, S.; Marchetti, F.; Di Bussolo, V. Stereoselective synthesis of new glycoconjugated ruthenium(II)-arene complexes as potential anticancer agents. In Proceedings of the EuroCarb XX, Leiden, The Netherlands, 30 June–4 July 2019. [Google Scholar]
  62. D’Amora, A.; Cucciolito, M.E.; Iannitti, R.; Morelli, G.; Palumbo, R.; Ruffo, F.; Tesauro, D. Pyridine ruthenium(III) complexes entrapped in liposomes with enhanced cytotoxic properties in PC-3 prostate cancer cells. J. Drug Deliv. Sci. Technol. 2019, 51, 552–558. [Google Scholar] [CrossRef]
  63. Valente, A.; Garcia, M.H.; Marques, F.; Miao, Y.; Rousseau, C.; Zinck, P. First polymer “ruthenium-cyclopentadienyl” complex as potential anticancer agent. J. Inorg. Biochem. 2013, 127, 79–81. [Google Scholar] [CrossRef]
  64. Lamač, M.; Horáček, M.; Červenková Šťastná, L.; Karban, J.; Sommerová, L.; Skoupilová, H.; Hrstka, R.; Pinkas, J. Harmless glucose-modified ruthenium complexes suppressing cell migration of highly invasive cancer cell lines. Appl. Organometal. Chem. 2019, 34, e5318. [Google Scholar] [CrossRef]
  65. Hanif, M.; Meier, S.M.; Nazarov, A.A.; Risse, J.; Legin, A.; Casini, A.; Jakupec, M.A.; Keppler, B.K.; Hartinger, C.G. Influence of the π-coordinated arene on the anticancer activity of ruthenium(II)carbohydrate organometallic complexes. Front. Chem. 2013, 1, 27. [Google Scholar] [CrossRef] [Green Version]
  66. Berger, I.; Hanif, M.; Nazarov, A.A.; Hartinger, C.G.; John, R.O.; Kuznetsov, M.L.; Groessl, M.; Schmitt, F.; Zava, O.; Biba, F.; et al. In vitro anticancer activity and biologically relevant metabolization of organometallic ruthenium com plexes with carbohydrate-based ligands. Chem. Eur. J. 2008, 14, 9046–9057. [Google Scholar] [CrossRef] [PubMed]
  67. Böge, M.; Fowelin, C.; Bednarski, P.; Heck, J. Diaminohexopyranosides as ligands in half-sandwich ruthenium(II), rhodium(III), and iridium(III) complexes. Organometallics 2015, 34, 1507–1521. [Google Scholar] [CrossRef]
  68. Florindo, P.; Marques, I.J.; Nunes, C.D.; Fernandes, A.C. Synthesis, characterization and cytotoxicity of cyclopentadienyl ruthenium(II) complexes containing carbohydrate-derived ligands. J. Organomet. Chem. 2014, 760, 240–247. [Google Scholar] [CrossRef]
  69. Florindo, P.R.; Pereira, D.M.; Borralho, P.M.; Rodrigues, C.M.P.; Piedade, M.F.M.; Fernandes, A.C. Cyclopentadienyl–ruthenium(II) and iron(II) organometallic compounds with carbohydrate derivative ligands as good colorectal anticancer agents. J. Med. Chem. 2015, 58, 4339–4347. [Google Scholar] [CrossRef]
  70. Hamala, V.; Martišová, A.; Červenková Šťastná, L.; Karban, J.; Dančo, A.; Šimarek, A.; Lamač, M.; Horáček, M.; Kolářová, T.; Hrstka, R.; et al. Ruthenium tetrazene complexes bearing glucose moieties on their periphery: Synthesis, characterization, and in vitro cytotoxicity. Appl. Organomet. Chem. 2020, 34, e5896. [Google Scholar] [CrossRef]
  71. Kilpin, K.J.; Crot, S.; Riedel, T.; Kitchen, J.A.; Dyson, P.J. Ruthenium(II) and osmium(II) 1,2,3-triazolylidene organometallics: A preliminary investigation into the biological activity of ‘click’ carbene complexes. Dalton Trans. 2014, 43, 1443–1448. [Google Scholar] [CrossRef] [Green Version]
  72. Florindo, P.R.; Pereira, D.M.; Borralho, P.M.; Costa, P.J.; Piedade, M.F.M.; Rodrigues, C.M.P.; Fernandes, A.C. New [(η5-C5H5)Ru(N–N)(PPh3)][PF6] compounds: Colon anticancer activity and GLUT-mediated cellular uptake of carbohydrate-appended complexes. Dalton Trans. 2016, 45, 11926–11930. [Google Scholar] [CrossRef]
  73. Kacsir, I.; Sipos, A.; Ujlaki, G.; Buglyó, P.; Somsák, L.; Bai, P.; Bokor, É. Ruthenium half-sandwich type complexes with bidentate monosaccharide ligands show antineoplastic activity in ovarian cancer cell models through reactive oxygen species production. Int. J. Mol. Sci. 2021, 22, 10454. [Google Scholar] [CrossRef]
  74. Hartinger, C.G.; Nazarov, A.A.; Ashraf, S.M.; Dyson, P.J.; Keppler, B.K. Carbohydrate-metal complexes and their potential as anticancer agents. Curr. Med. Chem. 2008, 15, 2574–2591. [Google Scholar] [CrossRef]
  75. Wang, X.; Guo, Z. Targeting and delivery of platinum-based anticancer drugs. Chem. Soc. Rev. 2013, 42, 202–224. [Google Scholar] [CrossRef] [PubMed]
  76. Fu, J.; Yang, J.; Seeberger, P.H.; Yin, J. Glycoconjugates for glucose transporter-mediated cancer-specific targeting and treatment. Carbohydr. Res. 2020, 498, 108195. [Google Scholar] [CrossRef] [PubMed]
  77. Franconetti, A.; López, Ó.; Fernandez-Bolanos, J.G. Carbohydrates: Potential sweet tools against cancer. Curr. Med. Chem. 2020, 27, 1206–1242. [Google Scholar] [CrossRef]
  78. Roux, C.; Biot, C. Ferrocene-based antimalarials. Future Med. Chem. 2012, 4, 783–797. [Google Scholar] [CrossRef]
  79. Ferreira, C.L.; Ewart, C.B.; Barta, C.A.; Little, S.; Yardley, V.; Martins, C.; Polishchuk, E.; Smith, P.J.; Moss, J.R.; Merkel, M.; et al. Synthesis, structure, and biological activity of ferrocenyl carbohydrate conjugates. Inorg. Chem. 2006, 45, 8414–8422. [Google Scholar] [CrossRef]
  80. Reddy, A.; Sangenito, L.S.; Guedes, A.A.; Branquinha, M.H.; Kavanagh, K.; McGinley, J.; Dos Santos, A.L.S.; Velasco-Torrijos, T. Glycosylated metal chelators as anti-parasitic agents with tunable selectivity. Dalton Trans. 2017, 46, 5297–5307. [Google Scholar] [CrossRef] [Green Version]
  81. Trivedi, R.; Deepthi, S.B.; Giribabu, L.; Sridhar, B.; Sujitha, P.; Kumar, C.G.; Ramakrishna, K.V.S. Synthesis, crystal structure, electronic spectroscopy, electrochemistry and biological studies of ferrocene-carbohydrate conjugates. Eur. J. Inorg. Chem. 2012, 2012, 2267–2277. [Google Scholar] [CrossRef]
  82. Panaka, S.; Trivedi, R.; Jaipal, K.; Giribabu, L.; Sujitha, P.; Kumar, C.G.; Sridhar, B. Ferrocenyl chalcogeno (sugar) triazole conjugates: Synthesis, characterization and anticancer properties. J. Organomet. Chem. 2016, 813, 125–130. [Google Scholar] [CrossRef]
  83. Coppens, I.; Asai, T.; Tomavo, S. Chapter 8-Biochemistry and Metabolism of Toxoplasma gondii: Carbohydrates, Lipids and Nucleotides. In Toxoplasma Gondii. In the Model Apicomplexan-Perspectives and Methods, 2nd ed.; Weiss, L.M., Kim, K., Eds.; Academic Press: Cambridge, MA, USA, 2014; pp. 257–295. [Google Scholar] [CrossRef]
  84. Fleige, T.; Fischer, K.; Ferguson, D.J.; Gross, U.; Bohne, W. Carbohydrate metabolism in the Toxoplasma gondii apicoplast: Localization of three glycolytic isoenzymes, the single pyruvate dehydrogenase complex, and a plastid phosphate translocator. Eukaryot. Cell 2007, 6, 984–996. [Google Scholar] [CrossRef] [Green Version]
  85. Coppin, A.; Dzierszinski, F.; Legrand, S.; Mortuaire, M.; Ferguson, D.; Tomavo, S. Developmentally regulated biosynthesis of carbohydrate and storage polysaccharide during differentiation and tissue cyst formation in Toxoplasma gondii. Biochimie 2003, 85, 353–361. [Google Scholar] [CrossRef] [PubMed]
  86. Blume, M.; Rodriguez-Contreras, D.; Landfear, S.; Fleige, T.; Soldati-Favre, D.; Lucius, R.; Gupta, N. Host-derived glucose and its transporter in the obligate intracellular pathogen Toxoplasma gondii are dispensable by glutaminolysis. Proc. Natl. Acad. Sci. USA 2009, 106, 12998–13003. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Giannini, F.; Furrer, J.; Süss-Fink, G.; Clavel, C.M.; Dyson, P.J. Synthesis, characterization and in vitro anticancer activity of highly cytotoxic trithiolato diruthenium complexes of the type [(η6-p-MeC6H4iPr)2Ru22-SR1)22-SR2)]+ containing different thiolato bridges. J. Organomet. Chem. 2013, 744, 41–48. [Google Scholar] [CrossRef]
  88. Păunescu, E.; Boubaker, G.; Desiatkina, O.; Anghel, N.; Amdouni, Y.; Hemphill, A.; Furrer, J. The quest of the best—A SAR study of trithiolato-bridged dinuclear ruthenium(II)-arene compounds presenting antiparasitic properties. Eur. J. Med. Chem. 2021, 222, 113610. [Google Scholar] [CrossRef] [PubMed]
  89. Desiatkina, O.; Paunescu, E.; Mosching, M.; Anghel, N.; Boubaker, G.; Amdouni, Y.; Hemphill, A.; Furrer, J. Coumarin-tagged dinuclear trithiolato-bridged ruthenium(II)arene complexes: Photophysical properties and antiparasitic activity. ChemBioChem 2020, 21, 2818–2835. [Google Scholar] [CrossRef]
  90. Giannini, F.; Bartoloni, M.; Paul, L.E.H.; Süss-Fink, G.; Reymond, J.L.; Furrer, J. Cytotoxic peptide conjugates of dinuclear areneruthenium trithiolato complexes. MedChemComm 2015, 6, 347–350. [Google Scholar] [CrossRef]
  91. Stíbal, D.; Therrien, B.; Süss-Fink, G.; Nowak-Sliwinska, P.; Dyson, P.J.; Čermáková, E.; Řezáčová, M.; Tomšík, P. Chlorambucil conjugates of dinuclear p-cymene ruthenium trithiolato complexes: Synthesis, characterization and cytotoxicity study in vitro and in vivo. J. Biol. Inorg. Chem. 2016, 21, 443–452. [Google Scholar] [CrossRef]
  92. Desiatkina, O.; Johns, S.K.; Anghel, N.; Boubaker, G.; Hemphill, A.; Furrer, J.; Păunescu, E. Synthesis and antiparasitic activity of new conjugates-organic drugs tethered to trithiolato-bridged dinuclear ruthenium(II)-arene complexes. Inorganics 2021, 9, 59. [Google Scholar] [CrossRef]
  93. Desiatkina, O.; Mösching, M.; Anghel, N.; Boubaker, G.; Amdouni, Y.; Hemphill, A.; Furrer, J.; Păunescu, E. New nucleic base-tethered trithiolato-bridged dinuclear ruthenium(II)-arene compounds: Synthesis and antiparasitic activity. Molecules 2022, 27, 8173. [Google Scholar] [CrossRef]
  94. Studer, V.; Anghel, N.; Desiatkina, O.; Felder, T.; Boubaker, G.; Amdouni, Y.; Ramseier, J.; Hungerbuhler, M.; Kempf, C.; Heverhagen, J.T.; et al. Conjugates containing two and three trithiolato-bridged dinuclear ruthenium(II)-arene units as in vitro antiparasitic and anticancer agents. Pharmaceuticals 2020, 13, 471. [Google Scholar] [CrossRef]
  95. Renfrew, A.K.; Juillerat-Jeanneret, L.; Dyson, P.J. Adding diversity to ruthenium(II)–arene anticancer (RAPTA) compounds via click chemistry: The influence of hydrophobic chains. J. Organomet. Chem. 2011, 696, 772–779. [Google Scholar] [CrossRef]
  96. Mandal, S.; Das, R.; Gupta, P.; Mukhopadhyay, B. Synthesis of a sugar-functionalized iridium complex and its application as a fluorescent lectin sensor. Tetrahedron Lett. 2012, 53, 3915–3918. [Google Scholar] [CrossRef]
  97. Schmollinger, D.; Kraft, J.; Ewald, C.; Ziegler, T. Synthesis of ruthenium and palladium complexes from glycosylated 2,20-bipyridine and terpyridine ligands. Tetrahedron Lett. 2017, 58, 3643–3645. [Google Scholar] [CrossRef]
  98. Mede, T.; Jager, M.; Schubert, U.S. “Chemistry-on-the-complex”: Functional Ru(II) polypyridyl-type sensitizers as divergent building blocks. Chem. Soc. Rev. 2018, 47, 7577–7627. [Google Scholar] [CrossRef] [PubMed]
  99. Cisnetti, F.; Gibard, C.; Gautier, A. Post-functionalization of metal–NHC complexes: A useful toolbox for bioorganometallic chemistry (and beyond)? J. Organomet. Chem. 2015, 782, 22–30. [Google Scholar] [CrossRef]
  100. van Hilst, Q.V.C.; Lagesse, N.R.; Preston, D.; Crowley, J.D. Functional metal complexes from CuAAC “click” bidentate and tridentate pyridyl-1,2,3-triazole ligands. Dalton Trans. 2018, 47, 997–1002. [Google Scholar] [CrossRef]
  101. Casas-Solvas, J.M.; Ortiz-Salmeron, E.; Gimenez-Martinez, J.J.; Garcia-Fuentes, L.; Capitan-Vallvey, L.F.; Santoyo-Gonzalez, F.; Vargas-Berenguel, A. Ferrocene-carbohydrate conjugates as electrochemical probes for molecular recognition studies. Chem. Eur. J. 2009, 15, 710–725. [Google Scholar] [CrossRef]
  102. Casas-Solvas, J.M.; Vargas-Berenguel, A.; Capitan-Vallvey, L.F.; Santoyo-Gonzalez, F. Convenient methods for the synthesis of ferrocene-carbohydrate conjugates. Org. Lett. 2004, 6, 3687–3690. [Google Scholar] [CrossRef]
  103. Meldal, M.; Tornoe, C.W. Cu-catalyzed azide-alkyne cycloaddition. Chem. Rev. 2008, 108, 2952–3015. [Google Scholar] [CrossRef]
  104. Kolb, H.C.; Finn, M.G.; Sharpless, K.B. Click chemistry: Diverse chemical function from a few good reactions. Angew. Chem. Int. Ed. Engl. 2001, 40, 2004–2021. [Google Scholar] [CrossRef]
  105. Singh, M.S.; Chowdhury, S.; Koley, S. Advances of azide-alkyne cycloaddition-click chemistry over the recent decade. Tetrahedron 2016, 72, 5257–5283. [Google Scholar] [CrossRef]
  106. Ibao, A.F.; Gras, M.; Therrien, B.; Süss-Fink, G.; Zava, O.; Dyson, P.J. Thiolato-bridged arene-ruthenium complexes: Synthesis, molecular structure, reactivity, and anticancer activity of the dinuclear complexes [(arene)2Ru2(SR)2Cl2]. Eur. J. Inorg. Chem. 2012, 2012, 1531–1535. [Google Scholar] [CrossRef]
  107. Soli, E.D.; DeShong, P. Advances in glycosyl azide preparation via hypervalent silicates. J. Org. Chem. 1999, 64, 9724–9726. [Google Scholar] [CrossRef]
  108. Paterson, S.M.; Clark, J.; Stubbs, K.A.; Chirila, T.V.; Baker, M.V. Carbohydrate-based crosslinking agents: Potential use in hydrogels. J. Polym. Sci. A Polym. Chem. 2011, 49, 4312–4315. [Google Scholar] [CrossRef]
  109. Mereyala, H.B.; Gurrala, S.R. A highly diastereoselective, practical synthesis of allyl, propargyl 2,3,4,6-tetra-O-acetyl-β-d-gluco, β-d-galactopyranosides and allyl, propargyl heptaacetyl-β-d-lactosides. Carbohydr. Res. 1998, 307, 351–354. [Google Scholar] [CrossRef]
  110. Roy, B.; Dutta, S.; Choudhary, A.; Basak, A.; Dasgupta, S. Design, synthesis and RNase A inhibition activity of catechin and epicatechin and nucleobase chimeric molecules. Bioorg. Med. Chem. Lett. 2008, 18, 5411–5414. [Google Scholar] [CrossRef]
  111. Wang, X.; Zhu, M.; Gao, F.; Wei, W.; Qian, Y.; Liu, H.K.; Zhao, J. Imaging of a clickable anticancer iridium catalyst. J. Inorg. Biochem. 2018, 180, 179–185. [Google Scholar] [CrossRef]
  112. Zabarska, N.; Stumper, A.; Rau, S. CuAAC click reactions for the design of multifunctional luminescent ruthenium complexes. Dalton Trans. 2016, 45, 2338–2351. [Google Scholar] [CrossRef]
  113. Wu, Q.; Liu, L.Y.; Li, S.; Wang, F.X.; Li, J.; Qian, Y.; Su, Z.; Mao, Z.W.; Sadler, P.J.; Liu, H.K. Rigid dinuclear ruthenium-arene complexes showing strong DNA interactions. J. Inorg. Biochem. 2018, 189, 30–39. [Google Scholar] [CrossRef]
  114. Zhao, J.; Li, S.; Wang, X.; Xu, G.; Gou, S. Dinuclear organoruthenium complexes exhibiting antiproliferative activity through DNA damage and a Reactive-Oxygen-Species-mediated endoplasmic reticulum stress pathway. Inorg. Chem. 2019, 58, 2208–2217. [Google Scholar] [CrossRef]
  115. Murray, B.S.; Menin, L.; Scopelliti, R.; Dyson, P.J. Conformational control of anticancer activity: The application of arene-linked dinuclear ruthenium(II) organometallics. Chem. Sci. 2014, 5, 2536–2545. [Google Scholar] [CrossRef] [Green Version]
  116. Chen, H.; Parkinson, J.A.; Novakova, O.; Bella, J.; Wang, F.; Dawson, A.; Gould, R.; Parsons, S.; Brabec, V.; Sadler, P.J. Induced-fit recognition of DNA by organometallic complexes with dynamic stereogenic centers. Proc. Natl. Acad. Sci. USA 2003, 100, 14623–14628. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Wang, H.-Y.; Qian, Y.; Wang, F.-X.; Habtemariam, A.; Mao, Z.-W.; Sadler, P.J.; Liu, H.-K. Ruthenium(II)–arene metallacycles: Crystal structures, interaction with DNA, and cytotoxicity. Eur. J. Inorg. Chem. 2017, 2017, 1792–1799. [Google Scholar] [CrossRef] [Green Version]
  118. Pettinari, C.; Pettinari, R.; Xhaferai, N.; Giambastiani, G.; Rossin, A.; Bonfili, L.; Eleuteri, A.M.; Cuccioloni, M. Binuclear 3,3′,5,5′-tetramethyl-1H,H-4,4′-bipyrazole Ruthenium(II) complexes: Synthesis, characterization and biological studies. Inorg. Chim. Acta 2020, 513, 119902. [Google Scholar] [CrossRef]
  119. Giannini, F.; Furrer, J.; Ibao, A.F.; Suss-Fink, G.; Therrien, B.; Zava, O.; Baquie, M.; Dyson, P.J.; Stepnicka, P. Highly cytotoxic trithiophenolatodiruthenium complexes of the type [(η6-p-MeC6H4Pri)2Ru2(SC6H4-p-X)3]+: Synthesis, molecular structure, electrochemistry, cytotoxicity, and glutathione oxidation potential. J. Biol. Inorg. Chem. 2012, 17, 951–960. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Giannini, F.; Suss-Fink, G.; Furrer, J. Efficient oxidation of cysteine and glutathione catalyzed by a dinuclear areneruthenium trithiolato anticancer complex. Inorg. Chem. 2011, 50, 10552–10554. [Google Scholar] [CrossRef] [Green Version]
  121. Desiatkina, O.; Boubaker, G.; Anghel, N.; Amdouni, Y.; Hemphill, A.; Furrer, J.; Paunescu, E. Synthesis, photophysical properties and biological evaluation of new conjugates BODIPY–dinuclear trithiolato-bridged ruthenium(II)-arene complexes. ChemBioChem 2022, 23, e202200536. [Google Scholar] [CrossRef]
  122. Barna, F.; Debache, K.; Vock, C.A.; Kuster, T.; Hemphill, A. In vitro effects of novel ruthenium complexes in Neospora caninum and Toxoplasma gondii tachyzoites. Antimicrob. Agents Chemother. 2013, 57, 5747–5754. [Google Scholar] [CrossRef] [Green Version]
  123. McFadden, D.C.; Seeber, F.; Boothroyd, J.C. Use of Toxoplasma gondii expressing beta-galactosidase for colorimetric assessment of drug activity in vitro. Antimicrob. Agents Chemother. 1997, 41, 1849–1853. [Google Scholar] [CrossRef]
  124. Muller, J.; Aguado-Manser, A.; Manser, V.; Balmer, V.; Winzer, P.; Ritler, D.; Hostettler, I.; Arranz-Solis, D.; Ortega-Mora, L.; Hemphill, A. Buparvaquone is active against Neospora caninum in vitro and in experimentally infected mice. Int. J. Parasitol. Drugs Drug. Resist. 2015, 5, 16–25. [Google Scholar] [CrossRef] [Green Version]
  125. Fulmer, G.R.; Miller, A.J.M.; Sherden, N.H.; Gottlieb, H.E.; Nudelman, A.; Stoltz, B.M.; Bercaw, J.E.; Goldberg, K.I. NMR chemical shifts of trace impurities: Common laboratory solvents, organics, and gases in deuterated solvents relevant to the organometallic chemist. Organometallics 2010, 29, 2176–2179. [Google Scholar] [CrossRef] [Green Version]
  126. Ogawa, T.; Nakabayashi, S.; Shibata, S. Synthetic studies on nephritogenic glycosides. Synthesis of N-(β-l-Aspartyl)-α-d-glucopyranosylamine. Agric. Biol. Chem. 1983, 47, 281–285. [Google Scholar] [CrossRef]
Figure 1. Structures of dinuclear thiolato-bridged ruthenium(II)-arene complexes (AC) active against T. gondii, of various ruthenium(II)-arene complexes presenting carbohydrate functionalized ligands and anticancer activity (DJ), and of ferrocene-carbohydrate conjugates exhibiting antimalarial, antibacterial, and anticancer properties (KM).
Figure 1. Structures of dinuclear thiolato-bridged ruthenium(II)-arene complexes (AC) active against T. gondii, of various ruthenium(II)-arene complexes presenting carbohydrate functionalized ligands and anticancer activity (DJ), and of ferrocene-carbohydrate conjugates exhibiting antimalarial, antibacterial, and anticancer properties (KM).
Molecules 28 00902 g001
Scheme 1. Synthesis of the diruthenium intermediates bearing OH (2), NH2 (3) CH2CO2H (4), and CH2-OH (5) groups on one of the bridge thiols.
Scheme 1. Synthesis of the diruthenium intermediates bearing OH (2), NH2 (3) CH2CO2H (4), and CH2-OH (5) groups on one of the bridge thiols.
Molecules 28 00902 sch001
Scheme 2. Synthesis of the alkyne functionalized ester and amide diruthenium compounds 6, 7 and 8.
Scheme 2. Synthesis of the alkyne functionalized ester and amide diruthenium compounds 6, 7 and 8.
Molecules 28 00902 sch002
Scheme 3. Synthesis of the azide functionalized diruthenium complex 9.
Scheme 3. Synthesis of the azide functionalized diruthenium complex 9.
Molecules 28 00902 sch003
Scheme 4. Synthesis of the azido glucose derivative 10.
Scheme 4. Synthesis of the azido glucose derivative 10.
Molecules 28 00902 sch004
Scheme 5. Synthesis of the glucose and galactose azide 1416, and alkyne 17 and 18 derivatives (eq = equatorial, ax = axial).
Scheme 5. Synthesis of the glucose and galactose azide 1416, and alkyne 17 and 18 derivatives (eq = equatorial, ax = axial).
Molecules 28 00902 sch005
Scheme 6. Synthesis of the diruthenium glucose conjugates 19 and 20 from the alkyne ester and amide derivatives 6 and 7.
Scheme 6. Synthesis of the diruthenium glucose conjugates 19 and 20 from the alkyne ester and amide derivatives 6 and 7.
Molecules 28 00902 sch006
Scheme 7. Synthesis of the glucose 21, and galactose 22 and 23 diruthenium conjugates.
Scheme 7. Synthesis of the glucose 21, and galactose 22 and 23 diruthenium conjugates.
Molecules 28 00902 sch007
Scheme 8. Synthesis of the diruthenium glucose conjugate 24.
Scheme 8. Synthesis of the diruthenium glucose conjugate 24.
Molecules 28 00902 sch008
Scheme 9. Synthesis of the glucose and galactose diruthenium conjugates 25 and 26.
Scheme 9. Synthesis of the glucose and galactose diruthenium conjugates 25 and 26.
Molecules 28 00902 sch009
Figure 2. Clustered column chart showing the in vitro activities at 1 (A) and 0.1 (B) µM of the azide derivative 10 and of trithiolato diruthenium compounds 29 and 1926 on HFF viability and T. gondii β-gal proliferation. For each assay, standard deviations were calculated from triplicates and are displayed on the graph. Data for compounds 25 and 79 were previously reported [88,89,92,93].
Figure 2. Clustered column chart showing the in vitro activities at 1 (A) and 0.1 (B) µM of the azide derivative 10 and of trithiolato diruthenium compounds 29 and 1926 on HFF viability and T. gondii β-gal proliferation. For each assay, standard deviations were calculated from triplicates and are displayed on the graph. Data for compounds 25 and 79 were previously reported [88,89,92,93].
Molecules 28 00902 g002aMolecules 28 00902 g002b
Table 1. Results of the primary efficacy/cytotoxicity screening of the azide derivative 10 and of trithiolato diruthenium compounds 29 and 1926 in non-infected HFF cultures and T. gondii β-gal tachyzoites cultured in HFF. Non-infected HFF monolayers treated only with 0.1% DMSO exhibited 100% viability and 100% proliferation was attributed to T. gondii β-gal tachyzoites treated with 0.1% DMSO only. The compounds selected for determination of IC50 values against T. gondii β-gal are tagged with *. For each assay, standard deviations were calculated from triplicates.
Table 1. Results of the primary efficacy/cytotoxicity screening of the azide derivative 10 and of trithiolato diruthenium compounds 29 and 1926 in non-infected HFF cultures and T. gondii β-gal tachyzoites cultured in HFF. Non-infected HFF monolayers treated only with 0.1% DMSO exhibited 100% viability and 100% proliferation was attributed to T. gondii β-gal tachyzoites treated with 0.1% DMSO only. The compounds selected for determination of IC50 values against T. gondii β-gal are tagged with *. For each assay, standard deviations were calculated from triplicates.
Molecules 28 00902 i001
CompoundRHFF Viability (%)T. Gondii β-gal Growth (%)
0.1 µM1 µM0.1 µM1 µM
Diruthenium intermediates
2 a,*Molecules 28 00902 i00276 ± 646 ± 666 ± 142 ± 0
3 a,*Molecules 28 00902 i00374 ± 248 ± 157 ± 12 ± 0
4 aMolecules 28 00902 i00491 ± 473 ± 1114 ± 2110 ± 2
5 a,*Molecules 28 00902 i00580 ± 169 ± 62 ± 01 ± 0
Alkyne and azide functionalized diruthenium compounds
6 *Molecules 28 00902 i006101 ± 1100 ± 14 ± 00 ± 0
7a,*Molecules 28 00902 i007101 ± 096 ± 021 ± 20 ± 0
8 aMolecules 28 00902 i00871 ± 246 ± 652 ± 133 ± 1
9 a,*Molecules 28 00902 i00996 ± 164 ± 19 ± 11 ± 1
Azido glucose derivative
10Molecules 28 00902 i01098 ± 199 ± 0101 ± 199 ± 0
Diruthenium—glucose conjugates
19 *Molecules 28 00902 i011102 ± 197 ± 19 ± 10 ± 0
20 *Molecules 28 00902 i01298 ± 199 ± 0100 ± 11 ± 2
21 *Molecules 28 00902 i013100 ± 1101 ± 087 ± 11 ± 0
24 *Molecules 28 00902 i01499 ± 198 ± 066 ± 23 ± 0
25Molecules 28 00902 i01598 ± 197 ± 175 ± 227 ± 0
Diruthenium—galactose conjugates
22 *Molecules 28 00902 i016100 ± 0101 ± 073 ± 12 ± 0
23 *Molecules 28 00902 i01799 ± 1101 ± 185 ± 11 ± 0
26 *Molecules 28 00902 i01898 ± 199 ± 189 ± 10 ± 0
a Compounds reported previously [88,89,92,93].
Table 2. Half-maximal inhibitory concentration (IC50) values (µM) on T. gondii β-gal for 15 selected compounds and pyrimethamine (used as standard), and their effect at 2.5 µM on HFF viability.
Table 2. Half-maximal inhibitory concentration (IC50) values (µM) on T. gondii β-gal for 15 selected compounds and pyrimethamine (used as standard), and their effect at 2.5 µM on HFF viability.
Molecules 28 00902 i019
CompoundRIC50 T. Gondii β-gal (µM)[LS; LI] bSE cHFF Viability at 2.5 µM (%) dSD e
Pyrimethamine a 0.326[0.396; 0.288]0.051996
Diruthenium intermediates
2 aMolecules 28 00902 i0200.117[0.139; 0.098]0.051566
3 aMolecules 28 00902 i0210.153[0.185; 0.127]0.049515
5 aMolecules 28 00902 i0220.038[0.023; 0.060]0.11042
Alkyne and azide functionalized diruthenium compounds
6Molecules 28 00902 i0230.023[0.030; 0.018]0.50371
7 aMolecules 28 00902 i0240.038[0.050; 0.029]0.063341
9 aMolecules 28 00902 i0250.048[0.058; 0.040]0.139111
Diruthenium-glucose conjugates
19Molecules 28 00902 i0260.018[0.031; 0.011]0.387293
20Molecules 28 00902 i0270.110[0.151; 0.080]0.416771
21Molecules 28 00902 i0280.087[0.135; 0.056]0.27400
24Molecules 28 00902 i0290.328[0.437; 0.247]0.339662
Diruthenium-galactose conjugates
22Molecules 28 00902 i0300.298[0.364; 0.244]0.066731
23Molecules 28 00902 i0310.032[0.044; 0.023]0.278922
26Molecules 28 00902 i0320.153[0.178; 0.132]0.476972
a Data for pyrimethamine, and compounds 2, 3, 5, 7 and 9 were previously reported [88,89,92,93]. 2 and 3 did not fulfil the first screening selection criteria, but the IC50 values and viability of HFF at 2.5 µM were determined for comparison purpose. b Values at 95% confidence interval (CI); LS is the upper limit of CI and LI is the lower limit of CI. c The standard error of the regression (SE), represents the average distance that the observed values fall from the regression line. d Control HFF cells treated only with 0.25% DMSO exhibited 100% viability. e The standard deviation of the mean (three replicate experiments).
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Holzer, I.; Desiatkina, O.; Anghel, N.; Johns, S.K.; Boubaker, G.; Hemphill, A.; Furrer, J.; Păunescu, E. Synthesis and Antiparasitic Activity of New Trithiolato-Bridged Dinuclear Ruthenium(II)-arene-carbohydrate Conjugates. Molecules 2023, 28, 902. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules28020902

AMA Style

Holzer I, Desiatkina O, Anghel N, Johns SK, Boubaker G, Hemphill A, Furrer J, Păunescu E. Synthesis and Antiparasitic Activity of New Trithiolato-Bridged Dinuclear Ruthenium(II)-arene-carbohydrate Conjugates. Molecules. 2023; 28(2):902. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules28020902

Chicago/Turabian Style

Holzer, Isabelle, Oksana Desiatkina, Nicoleta Anghel, Serena K. Johns, Ghalia Boubaker, Andrew Hemphill, Julien Furrer, and Emilia Păunescu. 2023. "Synthesis and Antiparasitic Activity of New Trithiolato-Bridged Dinuclear Ruthenium(II)-arene-carbohydrate Conjugates" Molecules 28, no. 2: 902. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules28020902

Article Metrics

Back to TopTop