2,500
Views
181
CrossRef citations to date
0
Altmetric
Review

Ruthenium Complexes as Anticancer Agents: A Brief History and Perspectives

, ORCID Icon & ORCID Icon
Pages 5375-5392 | Published online: 03 Dec 2020

Abstract

Platinum (Pt)-based anticancer drugs such as cisplatin have been used to treat various cancers. However, they have some limitations including poor selectivity and toxicity towards normal cells and increasing chemoresistance. Therefore, there is a need for novel metallo-anticancers, which has not been met for decades. Since the initial introduction of ruthenium (Ru) polypyridyl complex, a number of attempts at structural evolution have been conducted to improve efficacy. Among them, half-sandwich Ru-arene complexes have been the most prominent as an anticancer platform. Such complexes have clearly shown superior anticancer profiles such as increased selectivity toward cancer cells and ameliorating toxicity against normal cells compared to existing Pt-based anticancers. Currently, several Ru complexes are under human clinical trials. For improvement in selectivity and toxicity associated with chemotherapy, Ru complexes as photodynamic therapy (PDT), and photoactivated chemotherapy (PACT), which can selectively activate prodrug moieties in a specific region, have also been investigated. With all these studies on these interesting entities, new metallo-anticancer drugs to at least partially replace existing Pt-based anticancers are anticipated. This review covers a brief description of Ru-based anticancer complexes and perspectives.

Introduction

Transition metal complexes consisting of organic ligands bound to the center metal have played an important role in terms of their applications related to human civilization.Citation1,Citation2 Among them, ruthenium (Ru) complexes have received attention in many aspects.Citation3,Citation4 Positioned in the center of the second row of the transition metal series, Ru shows both early and late transition metal properties.Citation5 Due to its Lewis acidic but less oxophilic nature, the element displays a distinct array of properties utilized in many industrial and scientific fields such as solar cells,Citation6 electronics,Citation7 alloys,Citation8 catalysts,Citation9,Citation10 and diagnostic and therapeutic agents.Citation11Citation16 Ru has also been studiedCitation17,Citation18 in the field of medicinal inorganic chemistry, which has constantly grown during the past few decades. Organometallic complexes are generally considered unstable in air or in wet conditions. However, a variety of bioactive Ru complexes, which are stable in aqueous and alcoholic solutions and less sensitive to oxygen and sulfur, have been developed.Citation19 Researchers from both academia and industry have been focusing on the development of noble Ru complexes with prominent bioactivity and bioavailability.

In physiological conditions, Ru ion is stable in two oxidation states, Ru(II) and Ru(III), and the reduced state is considered to be more reactive.Citation20 Both oxidation states accommodate a six-coordinated octahedral configuration, but it is possible that they could be coordinated by ligands with different geometries and are expected to participate in various biological redox reactions.Citation21 Proper variation of ancillary ligands can modulate the steric and electronic properties of a complex to enable the construction of a large platform of chiral Ru complexes. Labile axial ligands are expected to coordinate to the disease targets through ligand-exchange reactions with biomolecules. The rate of ligand exchange in Ru(II) complexes (ranging from 10−2 to 10−3 s−1) is similar to that of platinum (II), which is on the scale of an average cell’s lifetime.Citation22 Thus, Ru is considered to be an alternative to platinum (Pt)-based drugs. In particular, many Ru compounds are considered to be less toxic than Pt-based drugs, and some of them are quite selective for cancerous cells.Citation23 These phenomena might have arisen from the ability of Ru to mimic iron in binding to biomolecules.Citation24 Overexpressed transferrin receptors on cancer cells due to their increased demand for iron may efficiently deliver Ru complexes to cancer cells.Citation25

Since Dwyer et al first developed a series of bioactive Ru polypyridyl complexes 13 in 1952 (),Citation26 Ru has been a prominent subject in the search for therapeutic and diagnostic agents, and a number of bioactive Ru complexes have been reported.Citation27Citation31 The major research field is the synthesis of new Ru(II) and Ru(III) complexes as potential anticancer agents and the investigation of their mechanism of action.Citation32 Although the applications of metal drugs are mainly related to the treatment of cancer, a significant amount of research has also been conducted to obtain therapies for other uses such as antivirals,Citation33 antibiotics,Citation34,Citation35 and anti-parasitic agents.Citation36,Citation37 Ru complexes are expected to be effective against infections due to the same mechanisms as those in the treatment of cancer. Thus, most of the Ru compounds tested for their cytotoxicity in different tumor cells have also been assessed in terms of their antimicrobial activity. Another area of growth is the study of the interactions between DNA and Ru complexes owing to the recent expansion of their roles such as chemical and stereoselective probes of nucleic acid structures,Citation38 molecular light switching and bioimaging,Citation39 and DNA bioanalysis agents.Citation40 The structurally complex three-dimensional architectures of metal complexes are ideal templates for constructing DNA interaction systems. As a result, Ru complexes have received attention by virtue of their unique binding ability to DNA, together with their rich photophysical, photochemical, and electrochemical properties.

Figure 1 Ru complexes developed by Dwyer.Citation26

Figure 1 Ru complexes developed by Dwyer.Citation26

Carbon Monoxide-Releasing Ru Complexes

An emerging research field for Ru complexes is the preparation of Ru-based carbon monoxide (CO)-releasing scaffolds (CORMs) to provide novel vehicles for intracellular CO delivery.Citation41 Carbon monoxide (CO), which is produced endogenously from the heme oxygenase (HO)-catalyzed degradation of heme,Citation42 is an important gas signaling molecule that plays significant anti-inflammatory, anti-apoptotic, anti-proliferative, and cytoprotective roles at low concentration.Citation43 Thus, controlled intracellular CO delivery in a specific target cell is expected to modulate cellular functions, and the development of biomaterials that can deliver CO into target cells in a dose-dependent manner is an attractive approach to achieve therapeutic values. However, its toxicity at high concentrations and the challenges for specific delivery to target sites are limiting its facile application. To overcome these problems, a wide range of Ru-based CORMs have been prepared for controlled dose-release of CO at the target tissue. The most extensively investigated CORM is Ru(CO)3Cl(glycinate), termed as CORM-3,Citation44 which has been reported to show interesting biological properties, including vasodilatory, anti-inflammatory, antibacterial, anti-ischemic and anti-apoptotic effects in preclinical studies.Citation45Citation47 There have been several recent review papers that have followed the update of CORMs. Therefore, this interesting topic will not be discussed further. Since studies on the preparation and evaluation of bioactive Ru complexes have been reported and have been the subject of many comprehensive reviews,Citation1Citation4,Citation48Citation50 this review will specifically focus on a brief history and future perspectives in anticancer agent research.

Ru-Based Anticancer Agents in Clinical Trials

The clinical success of the platinum (Pt) anticancer drug cisplatin (4) is an excellent example of how to advance a serendipitous discovery to a pharmaceutical.Citation51,Citation52 At present, three Pt-based anticancer compounds, cisplatin (4),Citation53 carboplatin (5),Citation54 and oxaliplatin (6),Citation55 have been approved and are used worldwide in clinical practice ().Citation56,Citation57 However, despite their clinical successes as chemotherapeutics, Pt-based drugs have some limitations: they are not active against many common types of cancer, drug resistance is common, and they cause a deplorable range of side effects such as nerve damage, hair loss, and nausea.Citation58Citation60 In search of alternative metal-based anticancer agents, Ru compounds have turned out to be the most promising candidates.Citation1,Citation61

Figure 2 Approved Pt-based anticancer drugs.

Figure 2 Approved Pt-based anticancer drugs.

A number of Ru-based anticancer agents have been developed to date, yet none of them are in clinical use as anticancer drugs. Successful entries to clinical trials of NAMI-A (7),Citation62 KP1019 (8),Citation63 NKP1339 (9),Citation64 and TLD1443 (10)Citation65,Citation66 together with many reports on the promising in vitro and in vivo activities of other types of Ru complexes have caused Ru-based chemotherapeutics to be seen as a major area in anticancer drug research ().Citation67,Citation68 Despite their structural similarity, NAMI-A and KP1019 have shown quite different in vitro and in vivo activities. NAMI-A showed antiangiogenic and antimetastatic activities in secondary tumorsCitation69,Citation70 whereas KP1019 is active in a broad spectrum of primary tumors.Citation71,Citation72 NKP1339, a sodium salt version of KP1019, was initially developed as a precursor in the formulation of KP1019 but reevaluated as a clinical candidate owing to its higher aqueous solubility, which allows for the clinical application of large doses to patients.Citation73 TLD1433 (10) entered Phase I and phase 2a clinical trials for bladder cancer treatment with photodynamic therapy (PDT).Citation66

Figure 3 Anticancer Ru complexes in clinical trials.

Figure 3 Anticancer Ru complexes in clinical trials.

Development of Anticancer Half-Sandwich Arene-Ru Complexes

Perhaps the most prosperous structural moiety of Ru-based anticancer agents over the last few decades has been half-sandwich Ru-arene complexes containing 1,3,5-triaza-7-phosphatricyclo-[3.3.1.1]decane (PTA) ligand(s) called RAPTAs. Their pharmacological properties can be easily modulated by ligand modification (1118 as shown in ).Citation67,Citation74Citation76 Their structures are composed of a “piano stool” geometry where an η5 or 6 π-arene ligand forms the seat, and the combination of mono- and bidentate ligands forms three legs. The chelating nature of the bidentate ligand appears to be beneficial for anticancer activity. Ru-arene complexes can display both hydrophilic and hydrophobic properties, which are expected to exhibit not only additive but also synergistic effects. The robust Ru-arene unit, together with other finely tuned ligands, can create diverse structural deviation and various modes of interaction with biomolecular targets that provide a high potential for the development of anticancer drugs. The first stable monomeric benzene-Ru complexes were reported by Zelonka et al in 1972,Citation77 and the development of anticancer arene-Ru complexes was initiated by the observation of Tocher et al in 1992 that the cytotoxicity of metronidazole, an antibiotic agent, was increased when coordinating to a benzene-Ru dichloro complex.Citation78 Since then, the RAPTA family has been a focus of research, and a number of analogues that display in vitro and in vivo anticancer activities have been prepared and evaluated. Interestingly, RAPTA complexes displayed a similar spectrum of activity as NAMI-A in spite of their differences in oxidation state, ligands, charge, and geometry.Citation79 The target of arene-Ru anticancer compounds may be DNA or RNA, but serum proteins might also become targets.Citation80 The development of RAPTA analogues along with the evaluation of their bioactivity has been described in detail in previous reviews;Citation74Citation76 therefore, only recent advances in the development of RAPTA analogues will be discussed in this review.

Figure 4 Selected RAPTA complexes from Ang and Dyson.Citation75

Figure 4 Selected RAPTA complexes from Ang and Dyson.Citation75

Recent Advances in the Development of RAPTA Family Anticancer Agents

Kurzwernhart et alCitation81 reported on the preparation and evaluation of a series of Ru(cymene) complexes with bioactive flavonol ligands, which are considered as topoisomerase inhibitors for use as anticancer agents. Studies on their mode of action have indicated that they form covalent bonds with DNA showing only a minor impact on the cell cycle but inhibit CDK2 and topoisomerase IIα in vitro. A cytotoxicity study against a panel of human cancer cell lines displayed IC50 values in a low μM range, which were lower than those of the parent compounds, flavonols. Complexes with para- and meta-substituted phenyl ligands exhibited lower IC50 values than unsubstituted ones. The structure of the most promising compound (19) is depicted in . Cȏrte-Real et alCitation82 prepared three cyclopentadienyl-Ru(II) bipyridyl complexes, including TM34. These complexes inhibited lactate production and trans-plasma membrane electron transport activity and showed inhibitory cell growth activity. Their uptake was facilitated without loss of activity when they were conjugated with transferrin. The most active compound, TM34 (20), exhibited cytotoxic activity against human tumor cell lines, A2780 and MDA-MB-231, in low μM IC50 values. Pettinari et alCitation83 prepared other RAPTA complexes consisting of fixed acylpyrazolonato bidentate ligand with varying arenes and monodentate ligands. The antitumor activity of the complexes was shown to be highly dependent on the nature of the arene ligand, and the complexes with the hexamethylbenzene (hmb) group turned out to be the most effective. The antiproliferative activity of such complexes against four human cancer cell lines was determined, and three hmb-Ru complexes displayed dose- and cancer cell line-dependent IC50 values in the low μM range. In particular, the most promising compound [(hmb)Ru(Qbiph)(PTA)][PF6] (21) was active in all tumor cell lines with a potency comparable to the reported values of cisplatin. The hmb-Ru complexes might activate caspase activity, thereby inducing DNA fragmentation, accumulation of proapoptotic proteins, and down-regulation of antiapoptotic protein Bcl-2, which results in apoptosis.

Figure 5 Selected anticancer RAPTA complexes.

Figure 5 Selected anticancer RAPTA complexes.

A library of water-soluble 442 arene-Ru Schiff base (RAS) complexes was efficiently constructed via a one-pot combinatorial metal three-component reaction by Chow et al.Citation84 The screening of the library for the anticancer activity showed that several compounds had low μM IC50 values in the inhibition of cell viability against a panel of cancer cell lines. The most active compound, the RAS complex 4 (22), exhibited cytotoxicity superior to that of the positive control, cisplatin. A mechanistic study suggested compound 22 induced apoptosis via a p53-independent mechanism, suggesting that DNA is not its primary biological target. In addition, Yellol et alCitation85 synthesized a series of C, N−bidentate Ru(II) and Ir (III) complexes, [(cymene)RuCl(HL)] and [(Me5Cp)IrCl(HL)] having various 2-(4-varied phenyl)benzimidazole ligands. All the complexes displayed high cell growth inhibitory activity in tested cells including cisplatin-resistant A2780cisR. In general, the Ru cytotoxic activity of metal compounds was evaluated against a panel of cancer cell lines, and all compounds displayed high cell growth inhibitory activity in tested cells including cisplatin-resistant A2780cisR. In general, Ru complexes are more active than their corresponding iridium complexes, and substitution on the 4-position of the phenyl ring caused increases in the potency of both the Ru and Ir complexes. The most potent compound [(cym)RuCl(HL)(2-(4-ph)phenyl] (23) showed superior inhibitory activity compared to the positive control, cisplatin. Additionally, six novel RAPTA type complexes containing curcumin-based ligands and 1,3,5-triaza-7-phosphaadamantane (PTA) were synthesized and characterized by Pettinari et al.Citation86 The antitumor activity of the complexes was evaluated in vitro against human ovarian carcinoma cells, A2780 and A2780cisR, and a non-tumorous human embryonic kidney HEK293 cell. Despite leading to different hydrolysis rates, the presence of a methoxy substituent on the phenyl rings of curcumin did not strongly affect the biological activity. However, the PTA ligand significantly enhanced the activity and selectivity of Ru complexes compared to the previously reported values for the parent compound, curcumin. All the Ru-curcumin complexes showed superior cytotoxicity and a cancer cell selectivity index compared to the positive control, cisplatin. In particular, the most promising compound [(hmb)Ru(bdcure)(PTA)] (24) was approximately 70-fold more selective against cancer cells than a noncancerous HEK cell.Citation86 A (cym)Ru(5-bromo-8-hydroxyquinoline) (25) hampered cell proliferation, migration and invasion in a monolayer of cancer cells. In addition, it showed anti-metastatic activity in spheroid model.Citation87,Citation88

Recent Advances in the Development of Anticancer Ruthenium Polypyridyl Analogues

In search of Ru-based antitumor compounds with a different spectrum of activity and fewer side effects, scientists have continued to search traditional Ru polypyridyl complexes as potent anticancer agents.Citation4 Cadoso et alCitation89 reported the preparation of five water-soluble, near IR luminescent Ru(II) polypyridyl complexes. The complexes were bound to HSA protein via non-covalent interaction and were luminescent in the near IR region. This method allows visualization of cellular localization and the distribution of administrated metal complexes. The high uptake of the complexes into HCT116 cells was detected by luminescent confocal microscopy. The inhibition of the proliferation of the cancer cell lines, A2780, HCT116 p53+/+, and HCT116 p53−/− by the complexes was evaluated, and the IC50 values for all the complexes were comparable to that of cisplatin. Among them, RuphenImH (26) displayed the highest activity, and the hydrogen-bonding ability of the imidazole ligand seemed to play an important role in their cytotoxicity (). Similarly, Huang et alCitation90 synthesized an inert Ru(II) complex [Ru(bpy)(phpy)(dppz)]+ (27) as an effective anticancer agent. Compound 27 was found to be rapidly taken up by cancer cells so that ~90% of the complex was accumulated in the nuclei of cancer cells 2 h after incubation. The anticancer activity of 27 was screened against a panel of cancer cell lines, and the compound exhibited IC50 values with a range of 0.6–4.3 μM, which is an order of magnitude lower than that of cisplatin. The formation of a Ru-carbon covalent bond in 27 enhanced the stability and lipophilicity, which would be beneficial for penetration to the cancer cell nuclei. The high DNA binding affinity of 27 caused a disruption in the binding of the transcription factor NF-κB to DNA, thereby inhibiting cellular transcription and leading to irreversible cancer cell apoptosis. In addition, Zeng et alCitation91 reported Ru(II) anthraquinone complexes that were highly cytotoxic to both normoxic and hypoxic cancer cells. The complexes exhibit similar or superior cytotoxicity compared to cisplatin in HeLa, A549, and multidrug-resistant A549R tumor cell lines. Their anticancer activities were correlated to their lipophilicity and cellular uptake properties. The most active compound, 28, exhibited 46-fold and 61-fold higher cytotoxic potency than cisplatin in hypoxic cells and 3D multicellular tumor spheroids, respectively. Compound 28 was preferentially accumulated in the mitochondria of hypoxic HeLa cells and induced apoptosis through multiple synergistic pathways. Similarly, three Ru(II) complexes with a bidentate bisimidazolo ligand were synthesized and characterized by Xia et al.Citation92 An interaction study between human telomeric G-quadruplex DNA and Ru complexes showed that they tightly bind to the human telomeric DNA. Among them, [Ru(phen)2(biim)]2+ (29), was the most effective in the formation of mixed/hybrid type G-quadruplexes. Their antitumor activity was closely related to their ability to interact with G-quadruplex DNA so that 29 showed the highest inhibitory activities comparable to the positive control, cisplatin, against HeLa, A549, and HepG2 cells. Compound 29 can effectively promote the apoptosis of tumor cells by acting on mitochondrial apoptotic pathways. Additionally, Han et alCitation93 reported the preparation of Ru(II) polypyridyl complexes and the evaluation of their anticancer activities. All the compounds reduced the mitochondrial membrane potential and inhibited cell growth in A549 cells in the G0/G1 phase. The IC50 values of complexes against BEL-7402, A549, MG-63, and SK-BR-3 cells were in the low μM range, and the cytotoxicity of the most active compound, [Ru(phen)2(HDPIP)] (30), was comparable to that of cisplatin. Among the four cell lines, A549 was found to be the most sensitive, while MG-63 was the least sensitive. Compound 30 enhanced the level of reactive oxygen species (ROS) and decreased the mitochondrial membrane potential, suggesting a compound-induced apoptosis in A549 cells through an ROS-mediated mitochondrial dysfunction pathway. Additionally, Chen et alCitation94 synthesized three water-soluble Ru(II) complexes with chiral 4-(2,3-dihydroxypropyl)-formamide oxoaporphine (FOA) ligand and evaluated their in vitro and in vivo antitumor activities. The compounds effectively stabilized telomeric and G-quadruplex DNA in the promoter of c-myc, thereby acting as a telomerase inhibitor. In the in vitro cytotoxicity against six human tumor cell lines (BEL-7404, A549, MGC80-3, HeLa, Hep-G2, BEL-7402 and one normal liver cell line HL-7702), compound LC-003 (31) displayed the highest inhibitory activity that was comparable to the positive control, cisplatin. Compound 31 was more selective for the BEL-7404 tumor cell than the normal HL-7702 cell. Moreover, 31 exhibited in vivo inhibition efficacy on tumor growth in the BEL-7402 xenograft mouse model and a higher in vivo safety profile than cisplatin. Another series of polypyridyl compounds (32) showed comparable anticancer activity to that of cis-platin, especially for gastric cancer cells.Citation95,Citation96 and summarize anticancer activities of selected Ru complexes in in vitro and in/ex vivo system, respectively.

Table 1 In vitro Anticancer Activity of the Ru-Based Compounds Selected in the Literature

Table 2 In/ex vivo Anticancer Activity of the Ru-Based Compounds Selected in the Literature

Figure 6 Selected anticancer Ru polypyridyl complexes.

Figure 6 Selected anticancer Ru polypyridyl complexes.

Photoactivation of Ru Complexes

Poor selectivity and lack of efficacy have been the main bottlenecks for anticancer chemotherapies. Thus, the selective delivery of a photoactivatable agent to tumorous cells and its activation by irradiation is an attractive approach to enhance the cytotoxicity of tumor cells while minimizing side effects on normal cells.Citation99 Two types of photoactivation modalities for Ru-based anticancer agents have been developed: photodynamic therapy (PDT) and photoactivated chemotherapy (PACT). In PDT treatment,Citation100Citation103 photosensitizer molecules are administered either topically or intravenously, and the target tissue layers are irradiated with light of a specific wavelength after internalization of the agents into tumor cells.Citation17,Citation104 The excited photosensitizer molecules then activate nearby oxygen and/or biomolecules to generate a reactive oxygen species, mainly singlet oxygen (1O2), causing the death of cancer cells. Because light has to reach deeper tissue layers, the light of a long wavelength is usually chosen in PDT, and the two-photon excitation (TPE) is preferred as the photoactivation method.Citation105 The advantages of TPE are reduced scattering of near-infrared (NIR) photons in turbid biological tissues and better definition of the focal spot.Citation106 However, PDT is still hindered by poor depth efficacy and reduced toxicity against hypoxic cancer cells. Porfimer sodium (Photofrin),Citation107 aminolevulinic acid (ALA),Citation108 and methyl ester of ALA (Metvixia)Citation109 have been approved for clinical practice as photosensitizing agents by the FDA. In contrast, PACTCitation65,Citation110Citation112 utilizes light to induce activation of the internalized prodrug moiety independent of the presence of oxygen inside cells. Thus, PACT is an attractive approach to treat tumors in a hypoxic condition, which is the notorious characteristic of solid cancers.Citation113 The photoactivation might induce DNA cross-linking, release of cytotoxic compounds from molecular carriers, or activation of prodrugs by ligand displacement. Because Ru complexes feature at least one coordination site in them that can be occupied by a labile ligand, construction of photo-dissociable Ru-based drug carriers or prodrugs is feasible.Citation114 During the last few decades, a number of photoactive Ru compounds that have exhibited multiple types of biologically relevant activity have been reported. There has also been a recent thorough review of Ru-based PDT and PACT agents.Citation115

For an update, Liu et al reported Ru(II) polypyridyl complexes as mitochondria that targeted two photon-absorbing PDT photosensitizers.Citation116 These complexes exhibited efficient singlet oxygen generation in methanol, significant two photon absorption (TPA) cross sections, and a substantial amount of mitochondrial accumulation. They are virtually nontoxic towards 3D HeLa multicellular spheroids (MCSs) in the dark but triggered cell death by a generation of singlet oxygen upon either one- or two-photon irradiation. The most promising complex, RuL4 (35) displayed IC50 values of 9.6 μM in one-photon and 1.9 μM in two-photon PDT, respectively, against 3D MCSs, which were lower than those of the positive control, cisplatin (). Likewise, Chen et alCitation117 developed Ru-arene complexes as potential candidates for dual PDT and PACT agents. The most active compound, [(cymene)Ru(dpb)(py)]2+ (36), absorbs long wavelength light and generated reactive 1O2, leading to photocleavage of DNA. In addition, compound 36 underwent photo-dissociation of both dpb and py ligands upon irradiation, and the resulting Ru-cymene fragment was bound to the nearby DNA bases. Hence, compound 36 has potential as a new type of antitumor agent with dual PACT and PDT pathways. The dissociated dpb ligand is a fluorescent process that might provide another opportunity to monitor real-time imaging of the photo-activated interaction between the biomolecule and Ru complex. Hence, compound 36 has potential as a new type of anticancer agent with dual PACT and PDT mechanisms. A dissociated dpb ligand can be a fluorescent probe that might provide another opportunity to monitor real-time images of photo-activated interaction between biomolecules and the Ru complex. Compound 36 showed significant light-enhanced cytotoxicity: IC50 of 27.6 μM in the dark and 4.0 μM under illumination against human lung carcinoma A549 cells. In addition, Huang et alCitation118 synthesized several mixed ligand Ru(II) terpyridyl complexes as PDT photosensitizing agents and investigated their photocytotoxicity. These complexes exhibited red luminescence between 670 and 710 nm and functioned as photosensitizers by generating both singlet oxygen and reactive radical species. They were located in the nucleus and exhibited significant photocytotoxicity upon irradiation. The most active compound, 37, exhibited a remarkable phototoxicity index (PI) together with a lower IC50 value than that of the positive control, cisplatin, toward human hepatocellular carcinoma Bel7402 and HepG2 cells. Similarly, Frei et alCitation119 prepared two Ru(II) polypyridyl complexes, Ru(DIP)2(bdt) (39) and [Ru(dqpCO2Me)(ptpy)]2+ (40), as PDT photosensitizers. Their phototoxicity was measured against the human cervical cancer cell line HeLa, and both compounds showed remarkable phototoxicity. Complex 40 displayed low μM range phototoxicity but no significant toxicity in the dark. However, complex 39 displayed μM range toxicity in the dark and nm range phototoxicity upon irradiation with a PI of 80, which is more impressive than that of the clinically approved photosensitizers, ALA and porfimer sodium. Both 39 and 40 displayed lower phototoxicity and PI values in similar experimental settings.Citation120 Around 67% of complex 39 taken up by the cell was accumulated in mitochondria. Nevertheless, the cellular uptake of 40 was shown to be diffused in all cellular compartments with a slight preference to the nucleus. The phototoxicity of both complexes against the two bacterial strains, S. aureus and E. coli, was also tested. Compound 39 effectively reduced cell viability in the Gram (+) strain S. aureus whereas no toxicity was observed against the Gram (−) strain E. coli. On the other hand, 40 effectively reduced the cell viability in both S. aureus and E. coli. The phototoxic profile of 40 against bacteria is particularly promising as Gram (−) bacterial strains are reported to be less affected by PDT than Gram (+) bacteria. One of the excellent examples would be 41 which showed a promising phototoxic profile.Citation121 It displayed IC50 > 100 μM in dark without irradiation while its IC50 is less than 1 μM when irradiated against CT-26 colon carcinoma cells reaching safety index >100.

Figure 7 Selected Ru-based PDT agents.

Figure 7 Selected Ru-based PDT agents.

As for the recent examples of PACT, Karaoun et alCitation122 constructed Ru(II) complexes with either one or two imidazole-based antifungal agent econazole ligands for dual applications in cell imaging and PACT agents. Both complexes were stable and luminescent in the dark, yet the irradiation of green light induced the release of an econazole ligand and turn-off of the luminescence. Although both complexes showed enhanced cytotoxicity and photocytoxicity against a panel of tumor cell lines compared to the parent drug econazole nitrate, which is known to induce apoptosis, Ru(phen)2(Ec)2Cl2 (42), which has two econazole ligands, displayed a higher PI value than the Ru(II) complex with one econazole ligand (). Compound 42 acts as a prodrug of econazole and offers several advantages, such as improved aqueous solubility and stability, enhanced intracellular accumulation, reduced toxicity, and real-time imaging of drug release by the turn-off luminescence response over the parent drug. Joshi et alCitation123 prepared a photolabile DMNPB ester capped Ru(II) complex (43) as a prodrug of the cytotoxic complex [Ru(dppz)2(CppH)]2+ (44),Citation124 which is known to disrupt the mitochondrial function. The hydrolytic stability test of 43 in the PBS buffer (pH 7.2) demonstrated that about 7% of 43 was hydrolyzed to be converted to 44 after 24 h in a dark environment. Compound 43 displayed negligible toxicity toward two cancerous (HeLa and U2OS) and non-cancerous (MRC-5) cells after 48 h in the dark. However, the cytotoxic action of the prodrug 43 can be regained in living cells under light illumination (350 nm), reaching a similar level of cytotoxicity as the parent cytotoxic compound 44, which is comparable to that of cisplatin. In order to photoactive in deep hypoxic legion in the body, compound 45 was developed to be excited by near-infrared light (NIR) via two-photon irradiation to treat melanoma cancer. The compound was readily absorbed by melanoma spheroid and showed rapid cell death in the hypoxic region.Citation125

Figure 8 Selected Ru-based PACT agents.

Figure 8 Selected Ru-based PACT agents.

Mode of Actions

Protein Binding

After intravenous administration, NAMI-A has a stronger interaction with human serum albumin (HSA) than KP1019, although both compounds bind to HSA in a noncovalent manner.Citation126Citation131 The stability of their noncovalent interaction has been shown to correlate with the ability of ligands to interact with the hydrophobic binding domains of HSA, and their different binding modes may play an important role in their distinct pharmacologic properties and efficacies. As shown in , which depicts the structure of the HSA-myristate-KP1019 complex,Citation132 Ru moieties bind to HAS. Both Ru metal centers are bound to the imidazole nitrogen of histidine 146 and histidine 242, which are located within the well-known drug binding sites, on subdomain IB (Ru binding site 1) and IIA (Ru binding site 2). The indazole ligands of KP1019 are important as binding sites recognizing moieties, which promote metal binding to other proteins in serum, which might lead to a decrease in selectivity and cytotoxicity and suggests an important reason for the pharmacologically different behavior between KP1019 and cisplatin, which was found to bind His residue located at the surface of HSA.Citation133

Figure 9 Crystal structure of the HAS-myristate-KP1019 complex. Two molecules of KP1019 bind to HAS at two histidine sites (H146 and H242).

Figure 9 Crystal structure of the HAS-myristate-KP1019 complex. Two molecules of KP1019 bind to HAS at two histidine sites (H146 and H242).

DNA Binding, Cytotoxicity, and Apoptosis

The cytotoxicity of cisplatin against cancer cells is mainly associated with binding to DNA via both interstrand and intrastrand cross-links, whereas the biologic targets of NAMI-A and KP1019 have not yet been totally elucidated.Citation134,Citation135 Both compounds are able to target DNA and proteins, implying the feasibility of either a different binding mode than that of cisplatin or the existence of multiple pathways. Both compounds are known as prodrugs; they are reduced to more reactive Ru(II) species by reducing molecules such as glutathione, cysteine, and ascorbic acid in a physiological medium.Citation136Citation138 There are two major proposed mechanisms by which Ru compounds are less toxic than platinum drugs in general: activation by reduction and the iron mimicking hypotheses. Activation by reduction theory is based on the observation that Ru(III) complexes are more inert than Ru(II), and cancerous cells, especially solid tumor issues, tend to have a greater reducing environment due to their lower oxygen level and pH condition than normal healthy cells.Citation139 Thus, the administrated Ru(III) compound causes minimal damage to healthy cells but can be activated to an Ru(II) oxidation state under a hypoxic environment inside cancerous cells.Citation140 The other postulation emerged from the fact that iron and Ru belong to the same group in the periodic table so that Ru is able to mimic iron during its interaction with biomolecules such as serum transferrin and albumin. Since rapidly growing tumor cells have a higher demand for iron uptake and often overexpress transferrin receptors on their cell surfaces, it is thought that they achieve only selective delivery and entry into metal complexes.Citation25 However, these theories have received considerable criticism. Even the question of how Ru compounds enter cells has been the subject of some debate in the literature.Citation141

NKP-1339, which is a leading clinical candidate among Ru-based anticancer compounds, showed very promising anticancer activity. As shown in ,Citation64 it decreased tumor volume and increased Td values (day when tumor volumes reach 300 mm3) in a xenograft model at a comparable level to sorafenib when administered i.v. once a week for two weeks. However, the tumor volume in both NKP-1339- and sorafenib-treated groups slowly caught up to that of the vehicle-treated group in the end. Combined with sorafenib, NKP-1339 reduced tumor volume more dramatically than its single treatment and kept it low for a long period of time. As for the mechanism of action, evidence to support other mechanisms than its direct interaction with DNA molecules was reported. It was proposed that if NKP-1339 is reduced to Ru(II) like other Ru-based anticancer compounds, then it could interact with unfolded protein response (UPR) machinery to regulate endoplasmic reticulum (ER) stress in cancer cells, or it could lead either to apoptosis or cell cycle arrest in cancer cells via mitochondrial damage or the p38 MAPK control ().

Figure 10 Anticancer activity of NKP-1339. (A) Hep3B xenografts in Balb/c SCID mice was treated with NKP-1339 (30 mg/kg, iv, once a week) and/or sorafenib (25 mg/kg, po, five consecutive days per week) for two weeks. (B) The mechanisms underlying the anticancer activity of representative Ru complexes: KP1019 and NKP-1339.

Figure 10 Anticancer activity of NKP-1339. (A) Hep3B xenografts in Balb/c SCID mice was treated with NKP-1339 (30 mg/kg, iv, once a week) and/or sorafenib (25 mg/kg, po, five consecutive days per week) for two weeks. (B) The mechanisms underlying the anticancer activity of representative Ru complexes: KP1019 and NKP-1339.

Conclusion and Perspectives

Over the last decades, Ru complexes have been targets of considerable attention and the fields of their application have rapidly grown. Ru complexes are in six-coordinated octahedral configurations, and proper variations of the ligands can allow the construction of a large platform of Ru compounds. The major research field of Ru complexes is the synthesis of potential anticancer agents among which the most prominent structural moiety of Ru-based anticancer agents has been half-sandwich Ru-arene complexes that display both hydrophilic and hydrophobic properties. A number of Ru complexes have shown superior anticancer profiles such as increased selectivity toward cancer cells and ameliorating toxicity against normal cells compared to existing Pt-based anticancer drugs. As a result, four Ru-based anticancer agents, NAMI-A (7), KP1019 (8), and NKP1339 (9), and TLD1433 (10) have entered clinical trials, but only two entities, NKP1339 (9) and TLD1433 (10) are still under investigation at this point.Citation142 One of the mainstream trends of oncology drugs is so-called targeted therapy where the agents are developed with a specific molecular target in hand. Ru-complexes have few target molecules known and therefore might be considered as non-specific. However, given that numerous oncology regimens still include non-specific Pt-based anticancer drugs, a new generation of metalloanticancers to overcome the existing drawbacks, such as poor selectivity for cancer cells and high toxicity against normal cells, would be an attractive alternative.

In order to overcome the poor selectivity and lack of toxicity associated with chemotherapy, PDT and PACT, which can selectively activate prodrug moieties in a specific region, have become a promising strategy. Studies in these applications have demonstrated feasibility as non-invasive and organelle-specific therapies such as mitochondrial-Citation116 and lysosome-targeting.Citation102 Despite very promising in vitro results of Ru-based PDT and PACT agents, insufficient in vivo studies have hampered full assessment of the feasibility of such compounds in a clinical context, which we believe researchers will have to now focus on.

An interesting approach to overcome the uptake, efficacy and biocompatibility issues related to Ru complexes will be a nanomaterial-conjugated PDT.Citation142 Another idea worthy of consideration would be a drug combination to synergize the efficacy without increasing the toxicity and drug resistanceCitation143 such as a combination with poly(ADP-ribose)polymerase (PARP) inhibitorsCitation144 to treat BRCA wild-type triple-negative breast cancer (TNBC). Recent studies are shedding new light on the anticancer potential of Ru complexes. Some have shown a promising characterization as an immune-modulating anticancer agent,Citation145 excellent redox potential,Citation97 strong topoisomerase inhibitor,Citation98 while others have shown encouraging results such as antimetastatic activity, tubulin formation inhibition, and high selectivity against cancer cells over normal cells.Citation146Citation148 Recent investigations cover structurally novel scaffolds including electron-deficient ruthenium complexesCitation149 and macromolecular ligands such as dendrimers.Citation150

In this review, we present a brief description of Ru-based anticancer complexes. With all the studies on these interesting entities, it is evident that new metalloanticancer drugs with improved efficacy and selectivity, and less toxicity compared to existing Pt-based anticancers should be seen in clinical use to provide new hope for cancer patients.

Disclosure

The authors report no conflicts of interest in this work.

Acknowledgments

This work was supported by research funds from Hanyang University (HY-2014-N). These authors contributed equally as the first author: Sang Yeul Lee and Chul Young Kim.

References

  • MaD, HeH, LeungK, et al. Bioactive luminescent transition-metal complexes for biomedical applications. Angew Chem Int Ed. 2013;52(30):7666–7682. doi:10.1002/anie.201208414
  • MediciaS, PeanaaM, NurchibV, et al. Noble metals in medicine: latest advances. Coordin Chem Rev. 2015;284:329–350. doi:10.1016/j.ccr.2014.08.002
  • DragutanH, DragutanV, DemonceauA. Editorial of special issue ruthenium complex: the expanding chemistry of the ruthenium complexes. Molecules. 2015;20:17244–17274. doi:10.3390/molecules20091724426393560
  • LevinaA, MitraA, LayP. Recent developments in ruthenium anticancer drugs. Metallomics. 2009;1(6):458–470. doi:10.1039/b904071d21305154
  • GrubbsR. Ruthenium. Chem Eng News. 2015;3(36):112–113. doi:10.1021/cen-v081n036.p112
  • KuangD, ItoS, WengerB, et al. High molar extinction coefficient heteroleptic ruthenium complexes for thin film dye-sensitized solar cells. J Am Chem Soc. 2006;128(12):4146–4154. doi:10.1021/ja058540p16551124
  • CornellA, SimonssD. Ruthenium dioxide as cathode material for hydrogen evolution in hydroxide and chlorate solutions. J Electrochem Soc. 1993;140(11):3123–3129. doi:10.1149/1.2220996
  • SchutzR. Ruthenium enhanced titanium alloys. Platin Metals Rev. 1996;40:54–61.
  • VougioukalakisG, GrubbsR. Ruthenium-based heterocyclic carbene-coordinated olefin metathesis catalysts †. Chem Rev. 2010;110(3):1746–1787. doi:10.1021/cr900242420000700
  • ArockiamP, BruneauC, DixneufP. Ruthenium(II)-catalyzed C–H bond activation and functionalization. Chem Rev. 2012;112(11):5879–5918. doi:10.1021/cr300153j22934619
  • ZhouX, ZhuD, LiaoY, et al. Synthesis, labeling and bioanalytical applications of a tris(2,2′-bipyridyl)ruthenium(II)-based electrochemiluminescence probe. Nat Protoc. 2014;9(5):1146–1159. doi:10.1038/nprot.2014.06024743421
  • ZhangW, ZhaoD, ZhangR, et al. A ruthenium(II) complex based turn-on electrochemiluminescence probe for the detection of nitric oxide. Analyst. 2011;136(9):1867–1872. doi:10.1039/c0an01003k21387076
  • Del MármolJ, FilevichO, EtcheniqueR. A ruthenium−rhodamine complex as an activatable fluorescent probe. Anal Chem. 2010;82(14):6259–6264. doi:10.1021/ac101212820583746
  • XuW, ZuoJ, WangL, et al. Dinuclear ruthenium(II) polypyridyl complexes as single and two-photon luminescence cellular imaging probes. Chem Commun. 2006;106(17):2123–2125. doi:10.1039/c3cc48916g
  • ShadeC, KennedyR, RougeJ, et al. Duplex-selective ruthenium-based DNA intercalators. Chem Eur J. 2015;21(31):10983–10987. doi:10.1002/chem.20150209526119581
  • CookN, KilpatrickK, SegatoriL, et al. Detection of α-synuclein amyloidogenic aggregates in vitro and in cells using light-switching dipyridophenazine ruthenium(II) complexes. J Am Chem Soc. 2012;134(51):20776–20782. doi:10.1021/ja310028723237404
  • MjosK, OrvigC. Metallodrugs in medicinal inorganic chemistry. Chem Rev. 2014;114(8):4540–4563. doi:10.1021/cr400460s24456146
  • RenfrewA. Transition metal complexes with bioactive ligands: mechanisms for elective ligand release and applications for drug delivery. Metallomics. 2014;6:1324–1335. doi:10.1039/C4MT00069B24850462
  • CrabtreeR. The Organometallic Chemistry of the Transition Metals. 6th ed. Hoboken, NJ, USA: John Wiley & Sons, Inc.; 2014.
  • Jabłońska-WawrzyckaA, RogalaP, MichałkiewiczS, et al. Ruthenium complexes in different oxidation states: synthesis, crystal structure, spectra and redox properties. Dalton Trans. 2013;42(17):6092–6101. doi:10.1039/c3dt32214a23381742
  • StrasserS, PumpE, FischerR, et al. On the chloride lability in electron-rich second-generation ruthenium benzylidene complexes. Monatsh Chem. 2015;146(7):1143–1151. doi:10.1007/s00706-015-1484-x
  • ReedijkJ. Metal-ligand exchange kinetics in platinum and ruthenium complexes. Platin Metals Rev. 2008;52(1):2–11. doi:10.1595/147106708X255987
  • GasserG, Metzler-NolteN. The potential of organometallic complexes in medicinal chemistry. Curr Opin Chem Biol. 2012;16(1–2):84–91. doi:10.1016/j.cbpa.2012.01.01322366385
  • BrabecV, NovakovaO. DNA binding mode of ruthenium complexes and relationship to tumor cell toxicity. Drug Resist Updat. 2006;9(3):111–122. doi:10.1016/j.drup.2006.05.00216790363
  • GatterK, BrownG, TrowbridgeI, et al. Transferrin receptors in human tissues: their distribution and possible clinical relevance. J Clin Pathol. 1983;36(5):539–545. doi:10.1136/jcp.36.5.5396302135
  • DwyerF, GyarfasE, RogersW, et al. Biological activity of complex ions. Nature. 1952;170(4318):190–191. doi:10.1038/170190a012982853
  • KilahN, MeggersE. Sixty years young: the diverse biological activities of metal polypyridyl complexes pioneered by Francis P. Dwyer. Aust J Chem. 2013;42(9):1325–1332. doi:10.1071/CH12275
  • HaraD, KomatsuH, SonA, et al. Water-soluble phosphorescent ruthenium complex with a fluorescent coumarin unit for ratiometric sensing of oxygen levels in living cells. Bioconjug Chem. 2006;9(4):645–649. doi:10.1021/acs.bioconjchem.5b00093
  • ShiS, GengX, ZhaoJ, et al. Interaction of [Ru(bpy)2(dppz)]2+ with human telomeric DNA: preferential binding to G-quadruplexes over i-motif. Biochimie. 2010;92(4):370–377. doi:10.1016/j.biochi.2010.01.00320096325
  • NovakovaO, KasparkovaJ, VranaO, et al. Correlation between cytotoxicity and DNA binding of polypyridyl ruthenium complexes. Biochemistry. 1995;34(38):12369–12378. doi:10.1021/bi00038a0347547981
  • KomorA, BartonJ. The path for metal complexes to a DNA target. Chem Commun. 2013;49(35):3617–3630. doi:10.1039/c3cc00177f
  • AllardyceC, DysonP. Ruthenium in medicine: current clinical uses and future prospects. Platin Metals Rev. 2001;45:62–69.
  • ChenL, ZhangX, ZhangC, et al. Dual-color fluorescence and homogeneous immunoassay for the determination of human enterovirus 71. Anal Chem. 2011;83(19):7316–7322. doi:10.1021/ac201129d21838323
  • LiF, HarryE, BottomleyA, et al. Dinuclear ruthenium(ii) antimicrobial agents that selectively target polysomes in vivo. Chem Sci. 2014;5(2):685–693. doi:10.1039/C3SC52166D
  • LiF, CollinsG, KeeneF. Ruthenium complexes as antimicrobial agents. Chem Soc Rev. 2015;44(8):2529–2542. doi:10.1039/C4CS00343H25724019
  • DonniciC, AraujoM, OliveiraH, et al. Ruthenium complexes endowed with potent anti-trypanosoma cruzi activity: synthesis, biological characterization and structure–activity relationships. Bioorg Med Chem. 2009;17(14):5038–5043. doi:10.1016/j.bmc.2009.05.07119539479
  • IniguezE, SanchezA, VasquezM, et al. Metal–drug synergy: new ruthenium(II) complexes of ketoconazole are highly active against leishmania major and trypanosoma cruzi and nontoxic to human or murine normal cells. J Biol Inorg Chem. 2013;18(7):779–790. doi:10.1007/s00775-013-1024-223881220
  • GillM, ThomasJ. Ruthenium(II) polypyridyl complexes and DNA–from structural probes to cellular imaging and therapeutics. Chem Soc Rev. 2012;41:3179–3192.22314926
  • FriedmanA, ChambronJ, SauvageJ, et al. A molecular light switch for DNA: Ru(bpy)2(dppz)2+. J Am Chem Soc. 1990;112(12):4960–4962. doi:10.1021/ja00168a052
  • ZhangS, DingY, WeiH. Ruthenium polypyridine complexes combined with oligonucleotides for bioanalysis: a review. Molecules. 2014;19(8):11933–11987. doi:10.3390/molecules19081193325116805
  • ForestiR, HammadJ, ClarkJ, et al. Vasoactive properties of CORM-3, a novel water-soluble carbon monoxide-releasing molecule. Br J Pharmacol. 2004;142(3):453–460. doi:10.1038/sj.bjp.070582515148243
  • TenhunenR, MarverH, SchmidR. The enzymatic conversion of heme to bilirubin by microsomal heme oxygenase. Proc Natl Acad Sci USA. 1968;61(2):748–755. doi:10.1073/pnas.61.2.7484386763
  • MannB. CO-releasing molecules: a personal view. Organometallics. 2012;31(16):5728–5735. doi:10.1021/om300364a
  • ClarkJ, NaughtonP, ShureyS, et al. Cardioprotective actions by a water-soluble carbon monoxide–releasing molecule. Circ Res. 2003;93(2):e2–e8. doi:10.1161/01.RES.0000084381.86567.0812842916
  • WangP, LiuH, ZhaoQ, et al. Syntheses and evaluation of drug-like properties of CO-releasing molecules containing ruthenium and group 6 metal. Eur J Med Chem. 2014;74:199–215. doi:10.1016/j.ejmech.2013.12.04124463436
  • InabaH, FujitaK, UenoT. Design of biomaterials for intracellular delivery of carbon monoxide. Biomater Sci. 2015;3(11):1423–1438. doi:10.1039/C5BM00210A26252321
  • NguyenD, BoyerC. Macromolecular and inorganic nanomaterials scaffolds for carbon monoxide delivery: recent developments and future trends. ACS Biomater Sci Eng. 2015;1(10):895–913. doi:10.1021/acsbiomaterials.5b00230
  • Troian-GautierL, MoucheronC. RutheniumII complexes bearing fused polycyclic ligands: from fundamental aspects to potential applications. Molecules. 2014;19(4):5028–5087. doi:10.3390/molecules1904502824759069
  • ValenteA, GarciaM. Syntheses of macromolecular ruthenium compounds: a new approach for the search of anticancer drugs. Inorganics. 2014;2(1):96–114. doi:10.3390/inorganics2010096
  • SharmaA, GangradeD, BakshiS, et al. Ruthenium complexes: potential candidate for anti-tumour activity. Int J Chem Tech Res. 2014;6:828–837.
  • FrickerS. Metal based drugs: from serendipity to design. Dalton Trans. 2007;36(43):4903–4917. doi:10.1039/b705551j
  • GalanskiM, JakupecM, KepplerB. Update of the preclinical situation of anticancer platinum complexes: novel design strategies and innovative analytical approaches. Curr Med Chem. 2005;12(18):2075–2094. doi:10.2174/092986705463762616101495
  • RosenbergB, VancampL, TroskoJ, et al. Platinum compounds: a new class of potent antitumour agents. Nature. 1969;222(5191):385–386. doi:10.1038/222385a05782119
  • EisenbergerM, HornedoJ, SilvaH, et al. Carboplatin (NSC-241-240): an active platinum analog for the treatment of squamous-cell carcinoma of the head and neck. J Clin Oncol. 1986;4(10):1506–1509. doi:10.1200/JCO.1986.4.10.15063531424
  • ExtraJ, EspieM, CalvoF, et al. Phase I study of oxaliplatin in patients with advanced cancer. Cancer Chemother Pharmacol. 1990;25:299–303.2295116
  • ReedijkJ. Platinum anticancer coordination compounds: study of DNA binding inspires new drug design. Eur J Inorg Chem. 2009;2009(10):1303–1312. doi:10.1002/ejic.200900054
  • WongE, GiandomenicoC. Current status of platinum-based antitumor drugs. Chem Rev. 1999;99:2451–2466.11749486
  • McWhinneyS, GoldbergR, McLeodH. Platinum neurotoxicity pharmacogenetics. Mol Cancer Ther. 2009;8(1):10–16. doi:10.1158/1535-7163.MCT-08-084019139108
  • KarasawaT, SteygerP. An integrated view of cisplatin-induced nephrotoxicity and ototoxicity. Toxicol Lett. 2015;237(3):219–227. doi:10.1016/j.toxlet.2015.06.01226101797
  • MiltenburgN, BoogerdW. Chemotherapy-induced neuropathy: a comprehensive survey. Cancer Treat Rev. 2014;40(7):872–882. doi:10.1016/j.ctrv.2014.04.00424830939
  • TruduF, AmatoF, VaňharaP, et al. Coordination compounds in cancer: past, present and perspectives. J Appl Biomed. 2015;13(2):79–103. doi:10.1016/j.jab.2015.03.003
  • Rademaker-LakhaiJ, van den BongardD, PluimD, BeijnenJH, SchellensJH. A Phase I and pharmacological study with imidazolium-trans-DMSO-imidazole-tetrachlororuthenate, a novel ruthenium anticancer agent. Clin Cancer Res. 2004;10(11):3717–3727. doi:10.1158/1078-0432.CCR-03-074615173078
  • HartingerC, JakupecaM, Zorbas-SeifriedaS, et al. KP1019, A new redox-active anticancer agent – preclinical development and results of a clinical phase I study in tumor patients. Chem Biodivers. 2008;5:2140–2150.18972504
  • TrondlR, HeffeterP, KowolC, et al. NKP-1339, the first ruthenium-based anticancer drug on the edge to clinical application. Chem Sci. 2014;5(8):2925–2932. doi:10.1039/C3SC53243G
  • SmithenDA, YinH, BehMHR, et al. Synthesis and photobiological activity of Ru(II) dyads derived from pyrrole-2-carboxylate thionoesters. Inorg Chem. 2017;56(7):4121–4132. doi:10.1021/acs.inorgchem.7b0007228301148
  • MonroS, ColonKL, YinH, et al. Transition metal complexes and photodynamic therapy from a tumor-centered approach: challenges, opportunities, and highlights from the development of TLD1433. Chem Rev. 2019;119(2):797–828. doi:10.1021/acs.chemrev.8b0021130295467
  • Uss-FinkG. Areneruthenium complexes as anticancer agents. Dalton Trans. 2010;39(7):1673–1688. doi:10.1039/B916860P20449402
  • KostovaI. Ruthenium complexes as anticancer agents. Curr Med Chem. 2006;13(9):1085–1107. doi:10.2174/09298670677636094116611086
  • SavaG, BergamoA, ZorzetbS, et al. Influence of chemical stability on the activity of the antimetastasis ruthenium compound NAMI-A. Eur J Cancer. 2002;38(3):427–435. doi:10.1016/S0959-8049(01)00389-611818210
  • SannaB, DebiddaM, PintusG, et al. The anti-metastatic agent imidazolium trans-imidazoledimethylsulfoxide-tetrachlororuthenate induces endothelial cell apoptosis by inhibiting the mitogen-activated protein kinase/extracellular signal-regulated kinase signaling pathway. Arch Biochem Biophys. 2002;403(2):209–218. doi:10.1016/S0003-9861(02)00218-712139970
  • KapitzaS, PongratzM, JakupecM, et al. Heterocyclic complexes of ruthenium(III) induce apoptosis in colorectal carcinoma cells. J Cancer Res Clin Oncol. 2005;131(2):101–110. doi:10.1007/s00432-004-0617-015503135
  • BergerM, GarzonF, KepplerB, et al. Efficacy of new ruthenium complexes against chemically induced autochthonous colorectal carcinoma in rats. Anticancer Res. 1989;9:761–765.2764521
  • ThompsonDS, WeissGJ, JonesSF, et al. NKP-1339: maximum tolerated dose defined for first-in-human GRP78 targeted agent. J Clin Oncol. 2012;30(15_suppl):abstract#3033. doi:10.1200/jco.2012.30.15_suppl.3033
  • BennettM. Recent advances in the chemistry of arene complexes of ruthenium(0) and ruthenium(II). Coordin Chem Rev. 1997;166:225–254. doi:10.1016/S0010-8545(97)00024-6
  • AngW, DysonP. Classical and non-classical ruthenium-based anticancer drugs: towards targeted chemotherapy. Eur J Inorg Chem. 2006;20:4003–4018.
  • AngW, CasiniA, SavaG, et al. Organometallic ruthenium-based antitumor compounds with novel modes of action. J Organomet Chem. 2011;696(5):989–998. doi:10.1016/j.jorganchem.2010.11.009
  • ZelonkaR, BairdM. Benzene complexes of ruthenium(II). Can J Chem. 1972;50(18):3063–3072. doi:10.1139/v72-486
  • DaleL, TocherJ, DysonT, et al. Studies on DNA damage and induction of SOS repair by novel multifunctional bioreducible compounds. A metronidazole adduct of a ruthenium-arene compound. Anti-Cancer Drug Des. 1992;7:3–14.
  • DysonP. Systematic design of a targeted organometallic antitumour drug in pre-clinical development. Chimia Int J Chem. 2007;61(11):698–703. doi:10.2533/chimia.2007.698
  • YanY, MelchartM, HabtemariamA, et al. Organometallic chemistry, biology and medicine: ruthenium arene anticancer complexes. Chem Commun. 2005;41(38):4764–4776. doi:10.1039/b508531b
  • KurzwernhartA, KandiollerW, BachlerS, et al. Structure–activity relationships of targeted Ru II (η6-p-cymene) anticancer complexes with flavonol-derived ligands. J Med Chem. 2012;55(23):10512–10522. doi:10.1021/jm301376a23134291
  • Corte-RealL, MendesF, CoimbraJ, et al. Anticancer activity of structurally related ruthenium(II) cyclopentadienyl complexes. J Biol Inorg Chem. 2014;19(6):853–867. doi:10.1007/s00775-014-1120-y24562604
  • PettinariR, PettinariC, MarchettiF, et al. Arene–ruthenium(II) acylpyrazolonato complexes: apoptosis-promoting effects on human cancer cells. J Med Chem. 2014;57(11):4532–4542. doi:10.1021/jm500458c24793593
  • ChowM, LiconaC, WongD, et al. Discovery and investigation of anticancer ruthenium–arene schiff-base complexes via water-promoted combinatorial three-component assembly. J Med Chem. 2014;57(14):6043–6059. doi:10.1021/jm500455p25023617
  • YellolJ, PérezS, BucetaA, et al. Novel C, N-cyclometalated benzimidazole ruthenium(II) and iridium(III) complexes as antitumor and antiangiogenic agents: a Structure–Activity Relationship Study. J Med Chem. 2015;58(18):7310–7327. doi:10.1021/acs.jmedchem.5b0119426313136
  • PettinariR, MarchettiF, CondelloF, et al. Ruthenium(II)–arene RAPTA type complexes containing curcumin and bisdemethoxycurcumin display potent and selective anticancer activity. Organometallics. 2014;33(14):3709–3715. doi:10.1021/om500317b
  • RuizMC, KljunJ, TurelI, et al. Comparative antitumor studies of organoruthenium complexes with 8-hydroxyquinolines on 2D and 3D cell models of bone, lung and breast cancer. Metallomics. 2019;11(3):666–675. doi:10.1039/C8MT00369F30839008
  • KljunJ, LeonIE, PersicŠ, et al. Synthesis and biological characterization of organoruthenium complexes with 8-hydroxyquinolines. J Inorg Biochem. 2018;186:187–196. doi:10.1016/j.jinorgbio.2018.05.00929960150
  • CardosoC, LimaM, CheleskiJ, et al. Luminescent ruthenium complexes for theranostic applications. J Med Chem. 2014;57:4906–4915.24831959
  • HuangH, ZhangP, YuB, et al. Targeting nucleus DNA with a cyclometalated dipyridophenazineruthenium(II) complex. J Med Chem. 2014;57(21):8971–8983. doi:10.1021/jm501095r25313823
  • ZengL, ChenY, HuangH, et al. Cyclometalated ruthenium(II) anthraquinone complexes exhibit strong anticancer activity in hypoxic tumor cells. Chem Eur J. 2015;21:15308–15319.26338207
  • XiaY, ChenQ, QinX, et al. Studies of ruthenium(ii)-2,2′-bisimidazole complexes on binding to G-quadruplex DNA and inducing apoptosis in HeLa cells. New J Chem. 2013;37(11):3706–3715. doi:10.1039/c3nj00542a
  • HanB, JiangG, WangJ, et al. The studies on bioactivity in vitro of ruthenium(ii) polypyridyl complexes towards human lung carcinoma A549 cells. RSC Adv. 2014;4(77):40899–40906. doi:10.1039/C4RA07102F
  • ChenZ, QinQ, QinJ, et al. Water-soluble ruthenium(II) complexes with chiral 4‑(2,3-dihydroxypropyl)-formamide oxoaporphine (FOA): in vitro and in vivo anticancer activity by stabilization of G‑quadruplex DNA, inhibition of telomerase activity, and induction of tumor cell apoptosis. J Med Chem. 2015;58:4771–4789.25988535
  • Ramirez-RiveraS, PizarroS, GallardoM, et al. Anticancer activity of two novel ruthenium compounds in gastric cancer cells. Life Sci. 2018;213:57–65. doi:10.1016/j.lfs.2018.10.02430326218
  • BabakMV, AngWH. Multinuclear organometallic ruthenium-arene complexes for cancer therapy. Met Ions Life Sci. 2018. doi:10.1515/9783110470734-012
  • NotaroA, FreiA, RubbianiR, et al. Ruthenium(II) complex containing a redox-active semiquinonate ligand as a potential chemotherapeutic agent: from synthesis to in vivo studies. J Med Chem. 2020;63(10):5568–5584. doi:10.1021/acs.jmedchem.0c0043132319768
  • XiongK, QianC, YuanY, et al. Necroptosis induced by ruthenium(II) complexes as dual catalytic inhibitors of topoisomerase I/II. Angew Chem Int Ed Engl. 2020;59(38):16631–16637. doi:10.1002/anie.20200608932533618
  • KnollJ, TurroC. Control and utilization of ruthenium and rhodium metal complex excited states for photoactivated cancer therapy. Coord Chem Rev. 2015;282–283:110–126. doi:10.1016/j.ccr.2014.05.018
  • LiuJ, ZhangC, ReesTW, et al. Harnessing ruthenium(II) as photodynamic agents: encouraging advances in cancer therapy. Coord Chem Rev. 2018;363:17–28. doi:10.1016/j.ccr.2018.03.002
  • ZengL, KuangS, LiG, et al. A GSH-activatable ruthenium(ii)-azo photosensitizer for two-photon photodynamic therapy. Chem Commun. 2017;53(12):1977–1980. doi:10.1039/C6CC10330H
  • HuangH, YuB, ZhangP, et al. Highly charged ruthenium(II) polypyridyl complexes as lysosome-localized photosensitizers for two-photon photodynamic therapy. Angew Chem Int Ed. 2015;54(47):14049–14052.
  • HessJ, HuangH, KaiserA, et al. Evaluation of the medicinal potential of two ruthenium(II) polypyridine complexes as one- and two-photon photodynamic therapy photosensitizers. Chem Eur J. 2017;23(41):9888–9896. doi:10.1002/chem.20170139228509422
  • DolmansD, FukumuraD, JainR. Photodynamic therapy for cancer. Nat Rev Cancer. 2003;3(5):380–387. doi:10.1038/nrc107112724736
  • BortG, GallavardinT, OgdenD, et al. From one-photon to two-photon probes: “caged” compounds, actuators, and photoswitches. Angew Chem Int Ed. 2013;52(17):4526–4537. doi:10.1002/anie.201204203
  • OgawaK, KobukeY. Recent advances in two-photon photodynamic therapy. Anticancer Agents Med Chem. 2008;8(3):269–279. doi:10.2174/18715200878396186018393786
  • BaskaranR, LeeJ, YangS-G. Clinical development of photodynamic agents and therapeutic applications. Biomater Res. 2018;22(1):25. doi:10.1186/s40824-018-0140-z30275968
  • WongS, CampbellB, MasseyB, et al. A phase I trial of aminolevulinic acid-photodynamic therapy for treatment of oral leukoplakia. Oral Oncol. 2013;49(9):970–976. doi:10.1016/j.oraloncology.2013.05.01123845699
  • WiegellS, Hæder SdalM, EriksenP, et al. Photodynamic therapy of actinic keratoses with 8% and 16% methyl aminolaevulinate and home-based daylight exposure: a double-blinded randomized clinical trial. Br J Dermatol. 2009;160(6):1308–1314. doi:10.1111/j.1365-2133.2009.09119.x19416257
  • van RixelVHS, MoolenaarGF, SieglerMA, et al. Controlling with light the interaction between trans -tetrapyridyl ruthenium complexes and an oligonucleotide. Dalton Trans. 2018;47(2):507–516. doi:10.1039/C7DT03613B29230469
  • LameijerLN, ErnstD, HopkinsSL, et al. A red-light-activated ruthenium-caged NAMPT inhibitor remains phototoxic in hypoxic cancer cells. Angew Chem Int Ed. 2017;56(38):11549–11553. doi:10.1002/anie.201703890
  • LiA, YadavR, WhiteJK, et al. Illuminating cytochrome P450 binding: ru(ii)-caged inhibitors of CYP17A1. Chem Commun. 2017;53(26):3673–3676. doi:10.1039/C7CC01459G
  • FarrerN, SalassaL, SadlerP. Photoactivated chemotherapy (PACT): the potential of excited-state d-block metals in medicine. Dalton Trans. 2009;38(48):10690–10701. doi:10.1039/b917753a
  • SmithN, SadlerP. Photoactivatable metal complexes: from theory to applications in biotechnology and medicine. Phil Trans R Soc. 2013;371(1995):20120519. doi:10.1098/rsta.2012.0519
  • MariC, PierrozV, FerraribS, et al. Combination of Ru(ii) complexes and light: new frontiers in cancer therapy. Chem Sci. 2015;6(5):2660–2686. doi:10.1039/C4SC03759F29308166
  • LiuJ, ChenY, LiG, et al. Ruthenium(II) polypyridyl complexes as mitochondria-targeted two-photon photodynamic anticancer agents. Biomaterials. 2015;56:140–153. doi:10.1016/j.biomaterials.2015.04.00225934287
  • ChenY, LeiW, JiangG, et al. Fusion of photodynamic therapy and photoactivated chemotherapy: a novel Ru(ii) arene complex with dual activities of photobinding and photocleavage toward DNA. Dalton Trans. 2014;43(41):15375–15384. doi:10.1039/C4DT01755B25188424
  • HuangH, ZhangP, YuB, et al. Synthesis, characterization and biological evaluation of mixed-ligand ruthenium(ii) complexes for photodynamic therapy. Dalton Trans. 2015;44(39):17335–17345. doi:10.1039/C5DT02081F26387554
  • FreiA, RubbianiR, TubafardS, et al. Synthesis, characterization, and biological evaluation of new Ru(II) polypyridyl photosensitizers for photodynamic therapy. J Med Chem. 2014;57(17):7280–7292. doi:10.1021/jm500566f25121347
  • DelaeyE, van LaarF, de VosD, et al. A comparative study of the photosensitizing characteristics of some cyanine dyes. J Photochem Photobiol B. 2000;55(1):27–36. doi:10.1016/S1011-1344(00)00021-X10877064
  • KargesJ, KuangS, MaschiettoF, et al. Rationally designed ruthenium complexes for 1- and 2-photon photodynamic therapy. Nat Commun. 2020;11:3262. doi:10.1038/s41467-020-16993-032591538
  • KaraounN, RenfrewA. A luminescent ruthenium(ii) complex for light-triggered drug release and live cell imaging. Chem Commun. 2015;51(74):14038–14041. doi:10.1039/C5CC05172J
  • JoshiT, PierrozV, MariC, et al. A bis(dipyridophenazine)(2-(2-pyridyl)pyrimidine-4-carboxylic acid)ruthenium(II) complex with anticancer action upon photodeprotection. Angew Chem Int Ed. 2014;53(11):2960–2963. doi:10.1002/anie.201309576
  • PierrozV, JoshiT, LeonidovaA, et al. Molecular and cellular characterization of the biological effects of ruthenium(II) complexes incorporating 2-pyridyl-2-pyrimidine-4-carboxylic acid. J Am Chem Soc. 2012;134(50):20376–20387. doi:10.1021/ja307288s23181418
  • RazaA, ArcherS, FairbanksSD, et al. A dinuclear ruthenium(II) complex excited by near-infrared light through two-photon absorption induces phototoxicity deep within hypoxic regions of melanoma cancer spheroids. J Am Chem Soc. 2020;142(10):4639–4647. doi:10.1021/jacs.9b1131332065521
  • WebbM, WalsbyC. Control of ligand-exchange processes and the oxidation state of the antimetastatic Ru(III) complex NAMI-A by interactions with human serum albumin. Dalton Trans. 2011;40(6):1322–1331. doi:10.1039/c0dt01168a21210063
  • CetinbasN, WebbM, DublandJ, et al. Serum-protein interactions with anticancer Ru(III) complexes KP1019 and KP418 characterized by EPR. J Biol Inorg Chem. 2010;15(2):131–145. doi:10.1007/s00775-009-0578-519707803
  • TimerbaevA, HartingerC, AleksenkoS, et al. Interactions of antitumor metallodrugs with serum proteins: advances in characterization using modern analytical methodology. Chem Rev. 2006;106(6):2224–2248. doi:10.1021/cr040704h16771448
  • MessoriL, VilchezF, VilaplanaR, et al. Binding of antitumor ruthenium(III) complexes to plasma proteins. Met Based Drugs. 2000;7(6):335–342. doi:10.1155/MBD.2000.33518475965
  • AitkenJ, AntonyS, WeekleyC, et al. Distinct cellular fates for KP1019 and NAMI-A determined by X-ray fluorescence imaging of single cells. Metallomics. 2012;4(10):1051–1056. doi:10.1039/c2mt20072d22907648
  • WebbM, ChardR, Al-JoboryY, et al. Pyridine analogues of the antimetastatic Ru(III) complex NAMI-A targeting non-covalent interactions with albumin. Inorg Chem. 2012;51(2):954–966. doi:10.1021/ic202029e22224431
  • BijelicA, TheinerS, KepplerBK, et al. X-ray structure analysis of indazolium trans- [tetrachlorobis(1 H -indazole)ruthenate(III)] (KP1019) bound to human serum albumin reveals two ruthenium binding sites and provides insights into the drug binding mechanism. J Med Chem. 2016;59(12):5894–5903. doi:10.1021/acs.jmedchem.6b0060027196130
  • FerraroG, MassaiL, MessoriL, et al. Cisplatin binding to human serum albumin: a Structural Study. Chem Commun. 2015;51(46):9436–9439. doi:10.1039/C5CC01751C
  • GroesslM, TsybinY, HartingerC, et al. Ruthenium versus platinum: interactions of anticancer metallodrugs with duplex oligonucleotides characterised by electrospray ionisation mass spectrometry. J Biol Inorg Chem. 2010;15(5):677–688. doi:10.1007/s00775-010-0635-020213306
  • GroesslM, ZavaO, DysonP. Cellular uptake and subcellular distribution of ruthenium-based metallodrugs under clinical investigation versus cisplatin. Metallomics. 2011;3:591–599. doi:10.1039/c0mt00101e21399784
  • SantosR, van EldikR, SilvaD. Kinetic and mechanistic studies on reactions of diruthenium(II, III) with biologically relevant reducing agents. Dalton Trans. 2013;42(48):16796–16805. doi:10.1039/c3dt51763b24081178
  • JakupecM, ReisnerE, EichingerA, et al. Redox-active antineoplastic ruthenium complexes with indazole: correlation of in vitro potency and reduction potential. J Med Chem. 2005;48(8):2831–2837. doi:10.1021/jm049074215828821
  • MillisK, WeaverK, RabensteinD. Oxidation/reduction potential of glutathione. J Org Chem. 1993;58(15):4144–4146. doi:10.1021/jo00067a060
  • SchlugaP, HartingerC, EggerA, et al. Redox behavior of tumor-inhibiting ruthenium(III) complexes and effects of physiological reductants on their binding to GMP. Dalton Trans. 2006;14:1796–1802.
  • TrédanO, GalmariniC, PatelK, et al. Drug resistance and the solid tumor microenvironment. J Natl Cancer Inst. 2007;99(19):1441. doi:10.1093/jnci/djm13517895480
  • BergamoA, SavaG. Ruthenium anticancer compounds: myths and realities of the emerging metal-based drugs. Dalton Trans. 2011;40:7817–7823.21629963
  • ZengL, GuptaP, ChenY, et al. The development of anticancer ruthenium(II) complexes: from single molecule compounds to nanomaterials. Chem Soc Rev. 2017;46:5771–5804.28654103
  • BerndsenRH, WeissA, AbdulUK, et al. Combination of ruthenium(II)-arene complex [Ru(η6-p-cymene)Cl2(pta)] (RAPTA-C) and the epidermal growth factor receptor inhibitor erlotinib results in efficient angiostatic and antitumor activity. Sci Rep. 2017;7(1):43005. doi:10.1038/srep4300528223694
  • YusohNA, LeongSW, ChiaSL, et al. Metallointercalator [Ru(dppz) 2 (PIP)] 2+ renders BRCA wild-type triple-negative breast cancer cells hypersensitive to PARP inhibition. ACS Chem Biol. 2020;15(2):378–387. doi:10.1021/acschembio.9b0084331898884
  • WernitznigD, KiakosK, Del FaveroG, et al. First-in-class ruthenium anticancer drug (KP1339/IT-139) induces an immunogenic cell death signature in colorectal spheroids in vitro. Metallomics. 2019;11(6):1044–1048. doi:10.1039/C9MT00051H30942231
  • QinQ-P, WangZ-F, HuangX-L, et al. High in vitro and in vivo tumor-selective novel ruthenium(II) complexes with 3-(2′-Benzimidazolyl)-7-fluoro-coumarin. ACS Med Chem Lett. 2019;10(6):936–940. doi:10.1021/acsmedchemlett.9b0009831223451
  • Mohamed SubarkhanMK, RenL, XieB, et al. Novel tetranuclear ruthenium(II) arene complexes showing potent cytotoxic and antimetastatic activity as well as low toxicity in vivo. Eur J Med Chem. 2019;10:246–256. doi:10.1016/j.ejmech.2019.06.061
  • AcharyaS, MajiMR, PurkaitK, et al. Synthesis, structure, stability, and inhibition of tubulin polymerization by Ru II – p -cymene complexes of trimethoxyaniline-based schiff bases. Inorg Chem. 2019;58(14):9213–9224. doi:10.1021/acs.inorgchem.9b0085331241921
  • Soldevila‐BarredaJJ, AzmanovaM, Pitto‐BarryA, et al. Preclinical anticancer activity of an electron-deficient organoruthenium(II) complex. Chemmedchem. 2020;15(11):982–987. doi:10.1002/cmdc.20200009632237195
  • Sanz Del OlmoN, Maroto-DiazM, QuintanaS, et al. Heterofunctional ruthenium(II) carbosilane dendrons, a new class of dendritic molecules to fight against prostate cancer. Eur J Med Chem. 2020;207:112695. doi:10.1016/j.ejmech.2020.11269532882608