Open access peer-reviewed chapter

Nanoformulation as a Tool for Improve the Pharmacological Profile of Platinum and Ruthenium Anticancer Drugs

Written By

Valentina Uivarosi, Rodica Olar and Mihaela Badea

Submitted: 21 October 2016 Reviewed: 03 March 2017 Published: 23 August 2017

DOI: 10.5772/intechopen.68306

From the Edited Volume

Descriptive Inorganic Chemistry Researches of Metal Compounds

Edited by Takashiro Akitsu

Chapter metrics overview

1,974 Chapter Downloads

View Full Metrics

Abstract

Cisplatin and analogs are used for the treatment of some type of cancers in combination with organic cytostatics. Also, two ruthenium (III) complexes are in clinical trials as anticancer drugs. In order to overcome toxicity and resistance associated with this therapy and/or enhance stability, a large variety of formulations based on organic, inorganic, or hybrid matrix were developed and tested both in vivo and in vitro. The best results were obtained for systems properly functionalized in order to enhance the metal content and/or to specific target the tumor tissue through overexpressed receptors.

Keywords

  • platinum
  • ruthenium
  • anticancer metal-based drugs
  • nanoformulations
  • conjugation

1. Introduction

Despite the use of metal compounds in empirical medicines since the ancient civilization time of Mesopotamia, Egypt, India, and China, the pharmacological bases of their therapeutic action were just began to be understood in the last 50 years [1].

A milestone in the development of inorganic medicinal chemistry was represented by the serendipitously discovery of the anticancer agent cisplatin (Platinol) [2], which opened the gate of extensive and rigorous research for anticancer metal-based drugs. Cisplatin quickly became a successful antitumor agent, but over time, its severe side effects and installation of resistance led to the orientation of research toward finding new cisplatin analogs. Thus, “the second-generation platinum drugs” (e.g., carboplatin) with improved toxicological profiles and “the third-generation drugs” (e.g., oxaliplatin) overcoming cisplatin resistance have been developed [3].

Having in view the systemic administration, the patients experienced severe symptoms since cisplatin and its analogs, carboplatin and oxaliplatin, were introduced in cancer therapy. Moreover, the intrinsic or acquired resistance and the fact that many cancers are insensitive to platinum-based drug therapy started an assiduous search for formulations that are able to deliver these drugs with reduced toxicity but with a similar or even enhanced cytotoxic profile [49].

A promising strategy able to overcome most of the above limitations consists in embedding either the original drug or a precursor in a proper matrix that is able to release a high amount of active species at target site. As result, several formulations based on organic, inorganic, or hybrid materials were designed. Among organic-based materials, a large variety of lipids, polymers, or mixed species were developed as platinum- and ruthenium-based drug carriers while magnetite, gold, graphene, and silica were studied as inorganic-based materials for the same purpose. Moreover, hybrid materials based on functionalized graphene, gold, iron oxides, silica, or polinuclear complexes and polysilsesquioxanes were studied in order to facilitate the delivery of these drugs [69].

Beyond improving solubility and reducing toxicity, a main challenge of these formulations was to increase their selectivity for tumor cells in order to achieve an optimum pharmacological profile. The first formulation developed by platinum-based drugs embedding through noncovalent interactions generated systems with a low loading capacity. A proper functionalization of the embedding matrix with Pt(II) drugs or Pt(IV)/Ru(III) prodrugs and/or with a responsive stimulus or a targeting moiety provided species with an increased cytotoxicity [69].

A large variety of encapsulation matrices and conjugations were developed, and formulations exhibit a promising cytotoxicity against either multidrug resistant or platinum insensitive cancer cells.

Advertisement

2. Anticancer metallodrugs

Apart from extensive research undertaken in the field of platinum complexes, other metals or other therapeutic strategies have attracted attention in order to reduce the side effects, to mitigate the resistance, and to achieve the oral administration.

The anticancer metallodrugs known at this time belong to three main classes:

  1. anticancer therapeutics

  2. therapeutic radiopharmaceuticals

  3. photochemotherapeutic metallodrugs [10].

Numerous chemotherapeutic metallodrugs developed in the last 4 decades are based on a large variety of metals: Pt, Ru, Au, Sn, Al, Ga, In, Ti [1116]. Among the metal-based compounds, complexes of platinum (Pt(II) and Pt(IV)), ruthenium (Ru(II) and Ru(III)), gold (Au(I) and Au(III)), and titanium (Ti(IV)) are the most studied [13].

Therapeutic radiopharmaceuticals include a β-emitting radionuclide (89Sr, 90Y, 153Sm, 213Bi) or a α-emitting radionuclide (223Ra). In general, α- and β-(electrons) emitters are used in radiotherapy, while β+ (positrons) and γ-emitters are used in radiodiagnosis [14].

Utilization of photochemotherapeutic metallodrugs is based on the photodynamic therapy (PDT). In PDT, a photosensitizing agent is delivered in tumor cells, which are activated with light, generating cytotoxic singlet oxygen. Starting to observation that Photofrin, a haematoporphyrin derivative is a strong chelator, forming a complex with ZnIIin vivo, some photochemotherapeutic metallodrugs have been developed [15].

The main platinum-based anticancer drugs currently used in clinic are presented in Table 1, while the emerging platinum- and ruthenium-based anticancer agents are listed in Table 2.

Table 1.

Platinum-based anticancer drugs currently used in clinic.

Table 2.

Platinum and ruthenium-based anticancer drugs subjected to clinical trials.

Advertisement

3. Platinum-based drugs nanoformulations

The clinical use of cisplatin and its analogs evidenced pharmacological deficiencies such as poor water solubility, low bioavailability, and short circulating time, besides toxicity and resistance. Moreover, a few types of cancers are sensitive to platinum-based drugs treatment.

Therefore, in the last decades, the researches were focused in designing drug delivery systems that are able to overcome these issues, but with preserving or even enhancing the drug efficacy. A brief overview concerning nanoscale drug delivery systems based on worldwide approved platinum-based cytostatic drugs cisplatin, carboplatin, and oxaliplatin is presented with focus on systems that advanced in clinical trials or exhibited promising pharmacological profile in vitro or in vivo preclinical assays.

3.1. Cisplatin-based nanoformulations

Cisplatin was the pioneering metallodrug introduced for the cancer treatment with the best result obtained in testicular cancer cure, for which a rate of 90% survival was achieved.

An impressive work was directed in the last time to overcome the severe side effects and intrinsic or acquired resistance by its inclusion in a proper matrix. This approach provided a way to extend its curative effect to other types of cancer proved so far to be insensitive to platinum-based drugs alone or in combination with other organic antineoplastic drugs. Its encapsulation into liposomes or polymeric species seems to provide the most promising formulations so far, since some of these formulations are currently in clinical trials.

Many formulations were developed by cisplatin encapsulation in the aqueous core of liposomes, with differences that consist in the composition of lipid bilayer, platinum content, and release profile. These attempts to incorporate cisplatin into liposomes were limited by its low both hydrophilicity and lipophilicity that resulted in a very low drug-lipid ratio and unstable systems, especially when injected into the blood stream [69].

In order to increase the liposomes stability, these systems were coated with a biocompatible hydrophilic polymer such as polyethylene glycol (PEG). Among these, lipoplatin was developed by cisplatin incorporation in a mixture of lipids from vegetable and animal sources, some being PEGylated [17]. An optimum pharmacological profile was observed in phase I clinical trial and significant improvements in patients with acquired resistance in phase II, in combination with gemcitabine, advanced this formulation in phase III clinical trials for both nonsmall-cell lung and pancreatic carcinoma [18]. Moreover, a preclinical study evidenced the potential of lipoplatin for cisplatin-resistant cervical cancer treatment [19].

A modest pharmacological profile was evidenced in clinical trials for a similar formulation SPI-77 as a result of low amount of cisplatin released [20, 21], while for LiPlaCis, a significant renal nephrotoxicity and infusion reactions were observed during phase I clinical trial [22].

As a result, the studies were directed to increase the amount of platinum species embedded either by using negatively charged phospholipids to entrap electrostatic [Pt(NH3)2(H2O)2]2+ species or by lipid bilayer functionalization and conjugation with platinum(II) or platinum(IV) species [7]. In this respect, some formulations with a high loading capacity were developed by cis-Pt(NH3)2and cis-Pt(NH3)2Cl moieties coordinated to carboxylate groups of lipids, and some exhibited a significant antitumor activity both in vitro and in vivo assays [23, 24].

Moreover, systems with a pendant group having selectivity for an overexpressed receptor in the cancer cells have been exploited to enhance the platinum species accumulation through receptor-mediated endocytosis. Such liposomal system targeting epidermal growth factor receptor (EGFR)-expressing tumors was developed by conjugation with sodium alginate and indeed exhibited enhanced delivery ability into ovarian tumor tissues and a reduced nephrotoxicity in mice [25].

A variety of polymeric formulations designed as micelle, hydrogels, nanoparticles, and nanocapsules were also studied as cisplatin carrier. The noncovalent encapsulation provided systems with similar or even lower efficacy in comparison with free cisplatin and as a result polymer conjugates were developed by Pt(II) or Pt(IV) species in reversible coordination to a functional group from the polymer backbone or its branches [69, 26].

Among these, nanoplatin (NC-6004) was obtained as micellar formulation by cisplatin entrapping in the core of polyethylene glycol-poly(glutamic acid) copolymer. The in vitro and in vivo preclinical assays evidenced a complete tumor regression as well as a low nephrotoxicity and neurotoxicity in C26 murine colon carcinoma cell [27]. The phase I trial evidenced a better tolerability and reduced side effects in comparison with cisplatin [28] and thus advanced nanoplatin in phase II trials for nonsmall-cell lung cancer, bladder cancer, and bile duct cancer, respectively [8].

The conjugated polymer (AP5280) was developed as nanoparticles by N-(2-hydroxypropyl) methacrylamide copolymer conjugation by cis-Pt(NH3)2 moiety to the peptidyl side chains (Gly-Phe-Leu-Gly) ended with amidomalonate group. This formulation exhibited an increased cytotoxicity in murine tumor models [29] and, moreover, evidenced reduced side effects in a phase I clinical trial conducted by intravenous infusion administration [30].

The conjugate designed by cis-Pt(NH3)2 moiety coordination to polyethylene glycol branched with citric acid exhibited an enhanced cytotoxicity in both sensitive and resistant HT1080 human fibro sarcoma cells, CT26 fibroblasts, and SKOV3 human ovarian cells [31], while another one based on poly(ethylene glycol)-poly(acrylic acid) copolymer and encapsulated in calcium phosphate evidenced its cytotoxicity against a lung cancer cisplatin-resistant cell line [32].

A good antitumor activity was also achieved by cis-Pt(NH3)2 moiety coordination to the carboxyl groups of poly(γ,L-glutamic acid)-based polymer [33], while by conjugation with polyamidoamines dendrimers developed nanocarriers that inhibit the subcutaneous B16F10 murine melanoma, a cisplatin insensitive tumor [34].

On the other hand, the conjugation and/or encapsulation of an organic cytostatic or a sensitive trigger together with platinum species were exploited to enhance the cytotoxicity of these formulations.

As a result, micellar carriers developed by poly(ethyleneglycol)-b-poly(L-glutamic acid)-b-poly(L-phenylalanine) tri-block copolymer conjugation with paclitaxel- and cisplatin-derived moieties exhibited an enhanced activity against A549 human lung tumor cells both in vitro and in vivo [35], while conjugates of both paclitaxel and cis,cis,trans-[Pt(NH3)2Cl2(OH)(HSucc)] (H2Succ: succinic acid) (Figure 1a) prodrug with poly(ethylene glycol)-b-poly(ε-caprolactone)-b-poly(l-lysine) tri-block amphiphilic biodegradable copolymer exhibited an enhanced efficacy in U14 cervical tumor line xenograft in mice as a result of the synergistic effect [36].

Figure 1.

Platinum (IV) complexes embedded into cisplatin based formulations: cis,cis,trans-[Pt(NH3)2Cl2(OH)(HSucc)] (a), cis,cis,trans-[Pt(NH3)2Cl2(HSucc)2] (b), and cis,cis,trans-[Pt(NH3)2Cl2(Bz)2] (c).

A combination of doxorubicin- and peptide-modified with cis-Pt(NH3)2 moiety loaded in positively charged mucoadhesive chitosan-polymethacrylic acid-based nanocapsules demonstrated an enhanced cytotoxicity against UMUC3 human urothelial carcinoma cell line [37]. Likewise, glutathione-sensitive micelles based on carboxymethyl chitosan crosslinked with 3,3′-dithiobis-N-hydroxysuccinimidyl propionate modified with folic acid exhibited synergistic cisplatin-doxorubicin effect against HeLa tumor cell line [38].

Another co-delivery system was developed by self-assembly of the anionic polyglutamic polymer cis-Pt(NH3)2 conjugated with an cationic metformin polymer. This formulation suppressed tumor growth for H460 human NSCLC xenografts in mice by a synergistic effect related to protein kinase α pathway activation and mammalian target rapamycin inhibition [39].

In order to achieve a high selectivity in targeting tumor cells, peptide and glycoside residues were inserted in the polymer backbone as groups that can be specifically recognized by the tumor tissue. This strategy resulted in thermosensitive nanoparticles obtained by cisplatin and indocyanine green loading in a complex matrix of poly(lactic-co-glycolic acid) copolymer and lipids functionalised with Gly-Cys-Gly-Ala-Ala-Asn-Leu heptapeptide. This formulation was designed to target MGC803 gastric tumor cells that overexpress the legumain and as a result exhibited a good activity in vitro [40].

Another formulation was developed as lyophilized system by cis-Pt(NH3)2(OH2) moiety coordination to carboxyl groups of hyaluronan, a naturally occurring glycosaminoglycan polysaccharide that targets tumor cells through specific interactions with CD44 receptor highly overexpressed in many cancers tissues. This conjugate demonstrated a suppressed cancer progression through intratracheal administration in Lewis lung carcinoma allografts in mice [41]. A platform targeting the same receptor was prepared by cisplatin incorporation in calcium phosphate and then embedded in hyaluronan-chitosan cross-linked polymer shell. These nanoparticles demonstrated target specific delivery in A549 human lung cancer cells confirmed by an eightfold increase of drug efficacy [42].

Some inorganic materials such as magnetite, graphene, gold, and silica were also studied in order to develop proper formulations for cisplatin delivery. The attempts to obtain nanoparticles based on these species have been discouraged by the low amount of cisplatin that can be noncovalent-retained and consequently promote an early release of the active species in the plasma. This problem has been solved either by coating the inorganic species-cisplatin assembly with an organic shell or by its surface functionalization with groups that are able to coordinate platinum species [69].

Following these strategies, an enhanced therapeutic effect in A549 human lung cancer xenograft model was obtained by magnetite-cisplatin assembly encapsulated in poly(vinyl alcohol) and poly(acrylic acid) [43]. Another formulation designed by cis-Pt(NH3)2 conjugation and magnetite embedded in (methacrylic acid)-g-poly(ethylene glycol methacrylate) polymer exhibited an enhanced anticancer efficacy in cisplatin-resistant HT-29 human colon adenocarcinoma model, particularly when a magnetic field gradient was applied at the tumor site [44].

An improved antitumor effect was also obtained either for gold nanoparticles PEGylated and cis,cis,trans-[Pt(NH3)2Cl2(HSucc)2] (Figure 1b) conjugated [45] or for that functionalized with oligonucleotide and cis,cis,trans-[Pt(NH3)2Cl2(OH)(HSucc))] conjugated [46]. It is to be pointed the higher cytotoxicity against cisplatin-resistant line exhibited by such formulations.

Nanoparticles developed by cis-Pt(NH3)2(OH) moiety coordination to functionalized mesoporous silica exhibited also an enhanced cytotoxicity on HT-29 colon cancer cell line [47].

Concerning graphene-based materials, a cisplatin nanotube conjugate modified with epidermal growth factor (EGF) proved an enhanced activity against EGF overexpressing head and neck squamous carcinoma cells [48], while functionalized multi-walled carbon nanotubes (MWCNTs) conjugated with cis,cis,trans-[Pt(NH3)2Cl2(Bz)2] (HBz: benzoyc acid) (Figure 1c) exhibited a selective accumulation in mice lungs [49].

Some hybrid materials based on coordination polymers were also developed as cisplatin carriers. Such supramolecular assembly was developed by [Tb2{Pt(NH3)2Cl2(Succ)2}3]n encapsulation in amorphous silica (Figure 2) as cytotoxic agent against HT-29 human colon carcinoma cell line [50].

Figure 2.

Hybrid nanoformulation developed by [Tb2{Pt(NH3)2Cl2(Succ)2}3] encapsulation.

Another platform was designed by hetero-metallic coordination polymer [Zn2{Pt(NH3)2Cl2(Ncp)2}]n (Ncp: N-carbamoyl phosphate) embedding in an asymmetric lipid layer modified with polyethylene glycol. This assembly, with a high amount of cisplatin incorporated, exhibited an enhanced efficacy in comparison with free cisplatin in H460 human nonsmall cell lung cancer and AsPC-1 human pancreatic cancer xenograft in mice [51].

A carrier system based on the same hetero-metallic coordination polymer and pyrolipid as photosensitizer exhibited a synergistic effect in cisplatin-resistant human head and neck cancer SQ20B xenograft in mice [52], while another formulation with small interfering RNA (siRNA) in addition and coated with a cationic lipid layer exhibited cytotoxicity both in vitro and in vivo against SKOV-3cisplatin-resistant ovarian cancer [53].

On the other hand, polysilsesquioxane-based hybrid nanomaterials developed by cis,cis,trans-[Pt(NH3)2Cl2(HptsSucc)2] (H2ptsSucc: propyltriethoxysilane succinic acid) polymerization (Figure 3) and coated with polyethylene glycol demonstrated an enhanced efficacy in combination with radiotherapy against A549 and H460 human lung cancer cells xenograft in mice [54].

Figure 3.

Hybrid nanoformulation developed by cis,cis,trans-[Pt(NH3)2Cl2(HptsSucc)2] polymerization.

These formulations can be internalized into the cancer tissues through passive or active transport. The passive transport is based on the ability of nanosystems to accumulate better in tumor tissue as a result of its increased permeability and poor lymphatic clearance, phenomenon known as enhanced permeability and retention (EPR) effect [55]. Moreover, the intratumoral nanoparticles content can be enhanced through an active transport facilitated by an overexpressed receptor.

Upon endo- or phagocytosis, the platinum species release is triggered in cytosol or other cellular compartments by several processes that can be acid, redox, and/or enzymatic assisted. For conjugated formulations, the cisplatin structure is restored either by reaction of Pt(II) species with chloride anions or by Pt(IV) species reduction with glutathione or ascorbic acid [69].

3.2. Carboplatin-based nanoformulations

The structural difference between cisplatin and carboplatin consists in replacing the chloride leaving groups by 1,1-ciclobutandicarboxylate as chelate ligand. Although carboplatin is often preferred over cisplatin in cancer therapy based on a lower nephrotoxicity, this exhibits a limited therapeutic efficacy related to its reduced uptake by the tumor cells. Moreover, the treatment induces myelosuppression and cross-resistance [14].

As a result, few studies were concerned on developing carboplatin-based formulations. Based on experience accumulated in cisplatin-based formulation development, carboplatin was embedded through noncovalent interactions especially in polymeric or hybrid materials, some proper functionalized in order to achieve either a targeted delivery or an enhanced efficacy, especially against multidrug resistant cancer cell lines.

Such a polymeric formulation was developed by loading in poly(D-L-lactide-co-glycolide) polymer. This nanocarrier exhibited an enhanced cellular uptake in both A549 lung and MA148 ovarian tumor cells [56], while that based on poly(ε-caprolactone) was also efficient uptakes and displayed a significant cytotoxicity in U-87 human glioma cell line, without inducing haemolysis [57]. Moreover, carboplatin-loaded apotransferrin and lactoferrin nanoparticles with high encapsulation efficacy exhibited a significantly cellular uptake and sustained intracellular drug retention in retinoblastoma cells [58], while a chitosan-based formulation demonstrated an enhanced antiproliferative effect against MCF-7 breast cancer cell line [59].

The hybrid materials were also studied in order to improve the pharmacological profile of carboplatin. Such supramolecular assembly based on multiple functionalizations of MWCNTs with amino groups resulted in a dramatic decrease of the MDA-MB-231 human mammary adenocarcinoma derived epithelial cells viability, which was related to superoxide anions production. This study also evidenced that expression of some proteins was inhibited, while the Beclin1 was overexpressed. As a result, most probably this system triggers the cell death through autophagy [60]. Another nanohybrid formulation developed by carboplatin loading in the nanographene oxide-gelatine material exhibited an enhanced efficacy in IMR-32 human neuroblastoma cell line [61].

3.3. Oxaliplatin-based nanoformulations

Oxaliplatin was introduced as first-line chemotherapeutic for the treatment of advanced colorectal cancer based on a different antineoplastic spectrum in comparison with cisplatin. However, the peripheral neuropathy and a moderate myelotoxicity in cumulative dose dependence were observed in many patients [62].

As a result, the attempts to improve its pharmacological profile and reduce the side effects resulted in several valuable formulations for this antineoplastic drug. Similar with cisplatin, a variety of organic, inorganic, and hybrid materials were studied for embedding either the original species [Pt(dach)(C2O4)] (dach: (1R,2R)-1,2-diaminocyclohexane) or another Pt(II) or Pt(IV) complex bearing dach as chelate ligand.

Among these, lipoxal developed as liposomal PEGylated formulation exhibited an acceptable pharmacological profile in a phase I clinical study for advanced gastrointestinal cancer [63]. By this formulation injected directly in F98 glioma implanted in rats, a reduced toxicity with preservation of the antitumor potential of oxaliplatin was achieved as well [64].

Similar with cisplatin, the efficacy of oxaliplatin-based formulation has been improved by surface of the liposomes modification with moieties that are able to assure either a specific targeting or a rapid release after the internalization of delivery system in tumor tissue.

These strategies resulted in developing a transferrin target sensitive liposomal formulation, which demonstrated increased tumor suppression in C-26 colon cancer cell line xenograft in mice as a result of transferrin receptor overexpression in this line [65]. This transferrin-targeted liposomal formulation is currently under phase II clinical investigation for the treatment of gastric cancer and gastroesophageal junction cancer [66].

By oxaliplatin encapsulation in PEGylated cationic liposomes, a formulation with a selective delivery in tumor vasculature was developed. The assays evidenced a complete suppressing tumor-induced angiogenesis and antitumor efficacy in mouse dorsal air sac as a result of dual-targeting both tumor cells and its vascular endothelial structure [67]. Moreover, the efficacy of this assembly can be improved by a sequential administration of oxaliplatin containing PEG-coated cationic liposomes [68].

Several oxaliplatin-based polymeric systems were also developed in order to enhance its cytotoxicity. Such oxaliplatin containing micelles (NC4016), in addition, proved the ability to overcome the oxaliplatin resistance in vivo are currently in clinical trials in patients with advanced solid tumors or lymphoma [69].

Another micellar formulation was developed by [Pt2(dach)2(dah)2](NO3)2 (dah: 1,2-diaminohexane) complex (Figure 4a) embedding into methoxylpoly(ethylene glycol)-b-poly(lactide-co-2-methyl-2-carboxylpropylene carbonate) (mPEG-b-P(LA-co-MCC)) copolymer. This pH sensitive assembly exhibited a significant cytotoxicity against H22 liver cancer cell line xenograft in mice [70].

Figure 4.

Platinum (IV) complexes embedded into oxaliplatin based formulations: [Pt2(dach)2(dah)2](NO3)2 (a), and [Pt(dach)(C2O4)(OH)(HSucc)] (b).

The polymeric systems were exploited not only to enhance the drug cytotoxicity through conjugation with Pt(dach) moieties, but for a combined delivery as well.

In this respect, micelles based on poly(ethylene glycol)-b-poly(glutamic acid) copolymer conjugated with Pt(dach) moiety demonstrated a potent tumor growth inhibition after an intraperitoneal injection in HeLa tumor cell xenograft in mice [71], while a similar micellar formulation inhibited the tumor growth in OCUM-2MLN scirrhous gastric cancer cell line and their lymphatic metastases in mice [72].

The polymer conjugate AP5346 was developed by Pt(dach) moiety coordination to the pH-sensitive amidomalonato chelating group from a N-(2-hydroxypropyl) methacrylamide-based copolymer structure. This conjugate exhibited an improved cytotoxicity in comparison with oxaliplatin in some colon tumor cell line xenograft in mice [73]. Based on pharmacological profile observed in patients with advanced solid tumors in phase I trial [69], this formulation advanced in phase II trial in recurrent ovarian cancer was initiated, but the results are so far disappointing [73].

Hybrid micelles containing mPEG-b-P(LA-co-MCC) copolymer conjugated with both Pt(dach) moiety and gemcitabine showed a low systemic toxicity and a synergic efficacy against MCF7 human breast cancer cell line xenograft in mice [74], while a similar system based on this copolymer conjugates with both [Pt(dach)(C2O4)(OH)(HSucc)] (Figure 4b) and daunorubicin showed reduced systemic toxicity and a synergistic effect in H22 hepatocarcinoma xenograft in mice [75].

In order to enhance the concentration of active species released in tumor tissue through a targeted delivery, some oxaliplatin-based polymer formulations were functionalized with glycoside residues and antibodies. Such polymeric nanoparticles were designed by carboplatin embedding in the supramolecular assembly of chitosan conjugated with hyaluronan and additionally coated with Eudragit S100. The oral administration of this formulation resulted in an enhanced activity in HT-29 cell line xenograft in mice [76].

Nanoparticles with a high amount of oxaliplatin embedded in a hybrid material consisting in a polymeric chitosan layer [77, 78] and a mixture of phospholipids conjugated with a thiolated antibody for tumour necrosis factor induced protein were developed as well [77]. Such formulations exhibited an increased cytotoxicity in comparison with oxaliplatin in HT-29 [77] and MCF7 cell lines [78].

Moreover, the functionalization allowed extending the cytotoxic effect to oxaliplatin insensitive tumors such as breast and gastric cancer. Thus, a pH-responsive nanocarrier was constructed by Pt(dach) moiety conjugation in citrate cross-linked chitosan matrix. The enhanced cytotoxicity of these nanoparticles in MCF-7 human breast cancer cell line was related to apoptosis induced in a caspase-dependent manner [67]. The nanogel system developed by embedding oxaliplatin in hydroxypropylcellulose-poly(acrylic acid) exhibited cytotoxicity against BGC823 human gastric cancer cell line [79].

Several systems based on hybrid materials were also developed for achieving an oxaliplatin enhanced delivery. Among these, superparamagnetic iron oxide nanoparticles encapsulated in pectin Ca2+ cross-linked exhibited 10-fold enhanced cytoxicity in comparison with free drug in MIA-PaCa-2 pancreas cancer cell line [80].

Another formulation developed by oxaliplatin incorporation into the inner cavity of PEGylated MWCNTs demonstrated a significantly improved cytotoxicity against HT-29 colorectal cell line [81], while similar nanocomposites additionally decorated with magnetite exhibited antitumor effect and low toxicity in HCT116 human colon cancer cell line xenograft in mice [82].

Naked gold nanoparticles functionalized with a thiolated poly(ethylene glycol) monolayer capped with a carboxylate group and conjugated with [Pt(dach)(H2O)2](NO3)2 yielded a supramolecular complex with about 280 Pt(dach) moieties per nanoparticle. This formulation demonstrated a similar or significant enhanced cytotoxicity in comparison with free oxaliplatin in A549 lung epithelial cancer cell line and HCT116, HCT15, HT29, and RKO colon cancer cell lines. Moreover, an unusual ability to penetrate the nucleus in the lung cancer cells was observed in these assays [83].

Mesoporous silica nanoparticles functionalised with carboxyl groups and conjugated with Pt(dach) moiety were also obtained with an improved cytotoxicity against HepG-2 human liver cell line [84].

Data concerning a platform constructed by [Zn2{Pt(dach)Cl2(Ncp)2}]n hetero-metallic coordination polymer conjugation to an asymmetric lipid bilayer modified with polyethylene glycol (Figure 5) were reported. This assembly with a high amount of platinum species incorporated exhibited cytotoxicity in H460 human nonsmall cell lung and AsPC-1 human pancreatic cancer cell lines xenograft in mice [53].

Figure 5.

Hybrid nanoformulation developed by [Zn2{Pt(dach)Cl2(Ncp)2}]n encapsulation.

Hybrid nanoparticles were also obtained by [Pt(dach)Cl2(triethoxysilylpropylsuccinate)2] base-catalyzed sol-gel polymerization similar to cisplatin derivative. Moreover, the silanol and carboxyl groups were functionalised with cyclic arginine-glycine-aspartate peptide and anisamide and then the surface was PEGylated. The cytotoxicity assay clearly indicated an increased uptake of this assembly by DLD-1 and HT-29 human adenocarcinoma cancer cells through integrin receptor and by AsPC-1 human pancreatic cancer cells through sigma receptor together with the tumor growth inhibition efficacy in pancreatic cancer xenograft in mice [85].

Advertisement

4. Ruthenium (III)-based drugs nanoformulations

The studies regarding ruthenium complexes as anticancer agents were developed as an alternative of platinum complexes, especially for their reduced toxicity, large spectrum of activities (including against cisplatin-resistant tumors) and selectivity [8688]. Among the various compounds of ruthenium investigated for their anticancer activity, two are in phase II clinical trials, namely NAMI-A (Table 2) as antimetastatic agent and KP1019 (Table 2) as antitumor for primary tumor site [8993].

Both are pseudo-octahedral complexes having four chloride ions in the equatorial plane. The axial ligands are imidazole and DMSO molecules in NAMI-A complex, while for KP1019 are two indazole molecules. Both complexes undergo hydrolysis in aqueous solutions (chloride ions being replaced by water and/or hydroxide ions) and interact with biological reductants leading to ruthenium (II) species. These two processes seem to provide the active species in the body [9496].

In order to improve the stability in aqueous systems, especially at physiological pH, and the delivery of drugs to the solid tumors, various drug delivery carriers have been designed and investigated. Two major ways were followed namely chemical conjugation and physical encapsulation [97].

4.1. Physical encapsulation of ruthenium-based drugs

Physical encapsulation is based on the capacity of carriers to retain the drug by physical bonds in a matrix. Different solid nanoparticles were used in order to encapsulate ruthenium complexes [97] such as poly(lactic acid) [98], mesoporous silica nanoparticles [99], or metal-organic frameworks [100]. The promising ruthenium (III) drug KP1019 was co-precipitated with poly(lactic acid) in a single oil-in-water emulsion with two different surfactants [98]. The obtained nanoparticles have an improved cytotoxicity comparing with KP1019.

4.2. Chemical conjugation of ruthenium-based drugs

4.2.1. Polymer conjugates

The main idea of this approach is to obtain a polymer, which contains a moiety that can act as ligand for ruthenium. In case of NAMI-A, this moiety can be an imidazole group. Thus, the Stenzel group [101] reports the polymerization of 4-vinil imidazole followed by addition of adequate ruthenium precursor complex. They obtained an amphiphilic co-polymer capable of self-assembly into micelles (Figure 6).

Figure 6.

NAMI-A conjugated to polymer.

The tests on ovarian and pancreatic cancer cells revealed a 1.5 times increased cytotoxicity for polymeric micelles. Furthermore, these were tested for antimetastatic activity on breast cancer cells proving a higher activity comparing to NAMI-A complex.

4.2.2. Lipid base conjutates/liposomes

The Paduano group focused on developing drug carriers for NANI-A analog, named AZIRu (Figure 7) [102108] and investigating their anticancer activity. Unlike NAMI-A, AZIRu contains a pyridine ligand instead of imidazole and sodium as counterion.

Figure 7.

AZIRu and selected amphiphilic nucleolipid-based AZIRu.

New amphiphilic derivatives of nucleosides have been developed in order to act as drug carriers for AZIRu complex. In detail, a nucleobase (thymidine or uridine), which was attached with a pyrimidilmethyl group at the N-3 position (in order to act as ligand for ruthenium) was selected as starting material. The resulted compounds were further bonded to one or two lipid residues (oleoyl or cholesteroxyacetyl) and one hydrophilic oligo(ethylene glycol) chain of variable lengths. There were thus obtained amphiphilic supramolecular aggregates, essentially liposomes [102105].

The nucleolipidic compounds proved to have similar instability in aqueous systems as NAMI-A and AZIRu, forming insoluble precipitates in few hours. In order to reduce the hydrolysis processes, the nucleolipidic compounds were formulated with biocompatible phospholipids, POPC (1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine) [103105] and DOTAP (1,2-dioleoyl-3-trimethylammoniumpropane) [106, 107]. The bioactivity of these RuIII-containing nucleolipids was tested on human and nonhuman cancer cells proving higher anticancer activity, higher stability in aqueous systems, and lower toxicity than AZIRu [108].

4.2.3. Dendrimers

The interest in dendrimers as drug carriers comes from their characteristics namely highly branched three-dimensional molecules containing functional groups at periphery, which can react with drug molecules. So far, only one potential anticancer ruthenium (III) drug, RAPTA-C, was incorporated into dendrimer (Figure 8) [109], but there is no study regarding the anticancer activity.

Figure 8.

Ruthenium (III) drug RAPTA-C and its dendrimeric nanoformulation.

Interactions of ruthenium (II) complexes with dendrimers and the anticancer activity of the resulted compounds, which are described in some reviews, have also attracted much interest [110, 111].

References

  1. 1. Sadler PJ, Muncie C, Shipman MA. Metals in Medicine in Biological Inorganic Chemistry, Structure and Reactivity. In: Bertini I, Gray HB, Stiefel EI, Selverstone Valentine J, editors. Sausalito, California: University Science Books; 2007. pp. 95-135
  2. 2. Rosenberg B, Vancamp L, Krigas T. Inhibition of cell division in Escherichia coli by electrolysis products from a platinum electrode. Nature. 1965;205:698-699. DOI:10.1038/205698a0
  3. 3. Kostova I. Platinum complexes as anticancer agents. Recent Patents on Anti-Cancer Drug Discovery. 2006;1:1-22. DOI:10.2174/157489206775246458
  4. 4. Mjos KD, Orvig C. Metallodrugs in medicinal inorganic chemistry. Chemical Reviews. 2014;114:4540-4563. DOI: dx.doi.org/10.1021/cr400460s
  5. 5. Kozarich JW. Medicinal Inorganic Chemistry: Promises and Challenges in Medicinal Inorganic Chemistry. In: Sessler JL, Doctrow SR, McMurry TJ, Lippard SJ, editors. Washington, DC: American Chemical Society;2005. pp. 4-14
  6. 6. Johnstone TC, Suntharalingam K, Lippard SJ. The next generation of platinum drugs: Targeted Pt(II) agents, nanoparticle delivery, and Pt(IV) prodrugs. Chemical Reviews. 2016;116:3436-3486. DOI: 10.1021/acs.chemrev.5b00597
  7. 7. Wani WA, Prashar S, Shreaz S, Gómez-Ruiz S. Nanostructured materials functionalized with metal complexes: In search of alternatives for administering anticancer metallodrugs. Coordination Chemistry Reviews. 2016;312:67-98. DOI: 10.1016/j.ccr.2016.01.001
  8. 8. Duan X, He C, Kron SJ, Lin W. Nanoparticle formulations of cisplatin for cancer therapy. WIREs Nanomedicine and Nanobiotechnolology. 2016;8:776-791. DOI: 10.1002/wnan.1390
  9. 9. Cheng Q, Liu Y. Multifunctional platinum-based nanoparticles for biomedical applications. WIREs Nanomedicine and Nanobiotechnolology. 2017;9:e1410. DOI: 10.1002/wnan.1410
  10. 10. Farrell N. Metal complexes as drugs and chemotherapeutic agents in volume 9: Applications of coordination chemistry. In: Ward MD, editor: from Comprehensive Coordination Chemistry II–From Biology to Nanotechnology. 2nd ed. editors: McCleverty JA, Meyer TJ. Amsterdam: Elsevier 2003, pp. 809-840
  11. 11. Muhammad N, Guo Z. Metal-based anticancer chemotherapeutic agents. Current Opinion in Chemical Biology. 2014;19:144-153. DOI: 10.1016/j.cbpa.2014.02.003
  12. 12. Gaynor D, Griffith DM. The prevalence of metal-based drugs as therapeutic or diagnostic agents: Beyond platinum. Dalton Transactions. 2012;41:13239-. DOI: 10.1039/c2dt31601c
  13. 13. Farrer NJ, Sadler PJ. Medicinal inorganic chemistry: State of the art, new trends, and a vision of the future. In: Alessio E, editors. Bioinorganic Medicinal Chemistry. Weinheim, Germany: Wiley-VCH Verlag & Co.;2011. pp. 1-48
  14. 14. Fricker SP. Metal based drugs: From serendipity to design. Dalton Transactions. 2007;43:4903-4917. DOI: 10.1039/b705551j
  15. 15. Hannon MJ. Metal-based anticancer drugs: From a past anchored in platinum chemistry to a post-genomic future of diverse chemistry and biology. Pure and Applied Chemistry. 2007;79:2243-2261. DOI:10.1351/pac200779122243
  16. 16. Barry NPE, Sadler PJ. Exploration of the medical periodic table: Towards new targets. Chemical Communications. 2013;49:5106-5131. DOI: 10.1039/c3cc41143e
  17. 17. Newman MS, Colbern GT, Working PK, Engbers C, Amantea MA. Comparative pharmacokinetics, tissue distribution, and therapeutic effectiveness of cisplatin encapsulated in long-circulating, pegylated liposomes (SPI-077) in tumor-bearing mice. Cancer Chemotherapy and Pharmacology. 1999;43:1-7. DOI: 10.1007/s002800050855
  18. 18. Boulikas T. Clinical overview on Lipoplatin™: A successful liposomal formulation of cisplatin. Expert Opinion on Investigational Drugs. 2009;18:1197-1218. DOI: 10.1517/13543780903114168.
  19. 19. Casagrande N, De Paoli M, Celegato M, Borghese C, Mongiat M, Colombatti A, Aldinucci D. Preclinical evaluation of a new liposomal formulation of cisplatin, lipoplatin, to treat cisplatin-resistant cervical cancer. Gynecologic Oncology 2013;131:744-752. DOI: 10.1016/j.ygyno.2013.08.041
  20. 20. White SC, Lorigan P, Margison GP, Margison JM, Martin F, Thatcher N, Anderson H, Ranson M. Phase II study of SPI-77 (sterically stabilized liposomal cisplatin) in advanced non-small-cell lung cancer. British Journal of Cancer. 2006;95:822-828. DOI:10.1038/sj.bjc.6603345
  21. 21. Seetharamu N, Kim E, Hochster H, Martin F, Muggia F. Phase II study of liposomal cisplatin (SPI-77) in platinum-sensitive recurrences of ovarian cancer. Anticancer Research. 2010;30:541-545
  22. 22. de Jonge MJ, Slingerland M, Loos WJ, Wiemer EA, Burger H, Mathijssen RH, Kroep JR, den Hollander MA, van der Biessen D, Lam M-H. Early cessation of the clinical development of LiPlaCis, a liposomal cisplatin formulation. European Journal of Cancer. 2010;46:3016-3021. DOI: 10.1016/j.ejca.2010.07.015
  23. 23. Guo S, Miao L, Wang Y, Huang L. Unmodified drug used as a material to construct nanoparticles: Delivery of cisplatin for enhanced anti-cancer therapy. Journal of Controlled Release. 2014;174:137-142. DOI: 10.1016/j.jconrel.2013.11.019
  24. 24. Li Q, Tian Y, Li D, Sun J, Shi D, Fang L, Gao Y, Liu H. The effect of lipocisplatin on cisplatin efficacy and nephrotoxicity in malignant breast cancer treatment. Biomaterials. 2014;35:6462-6472. DOI: 10.1016/j.biomaterials.2014.04.023
  25. 25. Wang Y, Zhou J, Qiu L, Wang X, Chen L, Liu T, Di W. Cisplatine-alginate conjugate liposomes for targeted delivery to EGFR-positive ovarian cancer cells. Biomaterials 2014;35:4297-4309. DOI: 10.1016/j.biomaterials.2014.01.035
  26. 26. Callaria M, Aldrich-Wright JR, de Souza PL, Stenzel MH. Polymers with platinum drugs and other macromolecular metal complexes for cancer treatment. Progress in Polymer Science2014;39:1614-1643. DOI: 10.1016/j.progpolymsci.2014.05.002
  27. 27. Uchino H, Matsumura Y, Negishi T, Koizumi F, Hayashi T, Honda T, Nishiyama N, Kataoka K, Naito S, Kakizoe T. Cisplatin-incorporating polymeric micelles (NC-6004) can reduce nephrotoxicity and neurotoxicity of cisplatin in rats. British Journal of Cancer 2005;93:678-687. DOI:10.1038/sj.bjc.6602772
  28. 28. Plummer R, Wilson R, Calvert H, Boddy A, Griffin M, Sludden J, Tilby M, Eatock M, Pearson D, Ottley C. A phase I clinical study of cisplatin incorporated polymeric micelles (NC-6004) in patients with solid tumours. British Journal of Cancer. 2011;104:593-598. DOI: 10.1038/bjc.2011.6
  29. 29. Lin X, Zhang Q, Rice J, Stewart D, Nowotnik D, Howell S. Improved targeting of platinum chemotherapeutics: The antitumour activity of the HPMA copolymer platinum agent AP5280 in murine tumour models. European Journal of Cancer. 2004;40:291-297. DOI: 10.1016/j.ejca.2003.09.022
  30. 30. Rademaker-Lakhai JM, Terret C, Howell SB,Baud CM, de Boer RF, Pluim D, Beijnen JH, Droz J-P. A phase I and pharmacological study of the platinum polymer AP5280 given as an intravenous infusion once every 3 weeks in patients with solid tumors. Clinical Cancer Research. 2004;10:3386-3395. DOI: 10.1158/1078-0432.CCR-03-0315
  31. 31. Haririan I, Alavidjeh MS, Khorramizadeh MR, Ardestani MS, Ghane ZZ, Namazi H. Anionic linear-globular dendrimer-cis-platinum(II) conjugates promote cytotoxicity in vitro against different cancer cell lines. International Journal of Nanomedicine. 2010;5:63-75. DOI: 10.2147/IJN.S8595
  32. 32. Ding Y, Zhai K, Pei P, Lin Y, Ma Y, Zhu H, Shao M, Yang X, Tao W. Encapsulation of cisplatin in a pegylated calcium phosphate nanoparticle (CPNP) for enhanced cytotoxicity to cancerous cells. Journal of Colloid and Interface Science. 2017;493:181-189. DOI: 10.1016/j.jcis.2017.01.032.
  33. 33. Xiong Y, Jiang W, Shen Y, Li H, Sun C, Ouahab A, Tu J. A poly(γ, L-glutamic acid)-citric acid based nanoconjugate for cisplatin delivery. Biomaterials. 2012;33:7182-7193. DOI: 10.1016/j.biomaterials.2012.06.071
  34. 34. Malik N, Evagorou EG, Duncan R. Dendrimer-platinate: A novel approach to cancer chemotherapy. Anti-Cancer Drugs. 1999;10:767-776
  35. 35. Song W, Tang Z, Li M, Lv S, Sun H, Deng M, Liu H, Chen X. Polypeptide-based combination of paclitaxel and cisplatin for enhanced chemotherapy efficacy and reduced side-effects. Acta Biomaterialia. 2014;10:1392-1402. DOI: 10.1016/j.actbio.2013.11.026
  36. 36. Xiao H, Song H, Yang Q, Cai H, Qi R, Yan L, Liu S, Zheng Y, Huang Y, Liu T, Jing X. A prodrug strategy to deliver cisplatin(IV) and paclitaxel in nanomicelles to improve efficacy and tolerance. Biomaterials. 2012;33:6507-6519. DOI: 10.1016/j.biomaterials.2012.05.049
  37. 37. Lu S, Xu L, Kang ET, Mahendran R, Chiong E, Neoh KG. Co-delivery of peptide-modified cisplatin and doxorubicin via mucoadhesive nanocapsules for potential synergistic intravesical chemotherapy of non-muscle-invasive bladder cancer. European Journal of Pharmaceutical Sciences. 2016;10;84:103-115. DOI: 10.1016/j.ejps.2016.01.013
  38. 38. Zhang X, Li L, Li C, Zheng H, Song H, Xiong F, Qiu T, Yang J. Cisplatin-crosslinked glutathione-sensitive micelles loaded with doxorubicin for combination and targeted therapy of tumors. Carbohydrate Polymers. 2017; 155:407-415. DOI: 10.1016/j.carbpol.2016.08.072
  39. 39. Xiong Y, Zhao Y, Miao L, Lin CM, Huang L. Co-delivery of polymeric metformin and cisplatin by self-assembled core-membrane nanoparticles to treat non-small cell lung cancer. Journal of Controlled Release. 2016;244(Pt A):63-73. DOI: 10.1016/j.jconrel.2016.11.005
  40. 40. Shi T, Gu L, Sun Y, Wang S, You C, Zhang X, Zhu J, Sun B. Enhanced legumain-recognition and NIR controlled released of cisplatin-indocyanine nanosphere against gastric carcinoma. European Journal of Pharmacology. 2017;5;794:184-192. DOI: 10.1016/j.ejphar.2016.11.039
  41. 41. Ishiguro S, Cai S, Uppalapati D, Turner K, Zhang T, Forrest WC, Forrest ML, Tamura M. Intratracheal administration of hyaluronan-cisplatin conjugate nanoparticles significantly attenuates lung cancer growth in mice. Pharmaceutical Research. 2016;33:2517-2529. DOI: 10.1007/s11095-016-1976-3
  42. 42. Suh MS, Shen J, Kuhn LT, Burgess DJ. Layer-by-layer nanoparticle platform for cancer active targeting. International Journal of Pharmaceutics. 2017;517:58-66. DOI: 10.1016/j.ijpharm.2016.12.006
  43. 43. Chiang C-S, Tseng Y-H, Liao B-J, Chen SY. Magnetically targeted nanocapsules for PAA-cisplatin-conjugated cores in PVA/SPIO shells via surfactant-free emulsion for reduced nephrotoxicity and enhanced lung cancer therapy. Advanced Healthcare Materials. 2015;4:1066-1075. DOI: 10.1002/adhm.201400794
  44. 44. Voulgari E, Bakandritsos A, Galtsidis S, Zoumpourlis V, Burke BP, Clemente GS, Cawthorne C, Archibald SJ, Tuček J, Zbořil R, Kantarelou V, Karydas AG, Avgoustakis K. Synthesis, characterization and in vivo evaluation of a magnetic cisplatin delivery nanosystem based on PMAA-graft-PEG copolymers. Journal of Controlled Release. 2016;243:342-356. DOI: 10.1016/j.jconrel.2016.10.021
  45. 45. Min Y, Mao C-Q, Chen S, Ma G, Wang J, Liu Y. Combating the drug resistance of cisplatin using a platinum prodrug based delivery system. Angewandte Chemie, International Edition. 2012;51:6742-6747. DOI: 10.1002/anie.201201562
  46. 46. Dhar S, Daniel WL, Giljohann DA, Mirkin CA, Lippard SJ. Polyvalent oligonucleotide gold nanoparticle conjugates as delivery vehicles for platinum(IV) warheads. Journal of the American Chemical Society. 2009;131:14652-14653. DOI: 10.1021/ja9071282
  47. 47. Gu J, Liu J, Li Y, Zhao W, Shi J. One-pot synthesis of mesoporous silica nanocarriers with tunable particle sizes and pendent carboxylic groups for cisplatin delivery. Langmuir. 2012;29:403-410. DOI.org/10.1016/j.ijpharm.2012.03.026
  48. 48. Bhirde AA, Patel V, Gavard J, Zhang G, Sousa AA, Masedunskas A, Leapman RD, Weigert R, Gutkind JS, Rusling JF. Targeted killing of cancer cells in vivo and in vitro with EGF-directed carbon nanotube-based drug delivery. ACS Nano. 2009;3:307-316. DOI: 10.1021/nn800551s
  49. 49. Li J, Pant A, Chin CF, Ang WH, Ménard-Moyon C, Nayak TR, Gibson D, Ramaprabhu S, Panczyk T, Bianco A, Pastorin G. In vivo biodistribution of platinum-based drugs encapsulated into multi-walled carbon nanotubes. Nanomedicine: Nanotechnology, Biology and Medicine. 2014;10:1465-1475. DOI: 10.1016/j.nano.2014.01.004
  50. 50. Rieter WJ, Pott KM, Taylor KML, Lin W. Nanoscale coordination polymers for platinum-based anticancer drug delivery. Journal of the American Chemical Society. 2008;130:11584-11585. DOI: 10.1021/ja803383k
  51. 51. Liu D, Poon C, Lu K, He C, Lin W. Self-assembled nanoscale coordination polymers with trigger release properties for effective anticancer therapy. Nature Communications. 2014;5. DOI:10.1038/ncomms5182
  52. 52. He C, Liu D, Lin W. Self-assembled core-shell nanoparticles for combined chemotherapy and photodynamic therapy of resistant head and neck cancers. ACS Nano. 2015;9:991-1003. DOI: 10.1021/nn506963h
  53. 53. He C, Liu D, Lin W. Self-assembled nanoscale coordination polymers carrying siRNAs and cisplatin for effective treatment of resistant ovarian cancer. Biomaterials. 2015;36:124-133. DOI:10.1016/j.biomaterials.2014.09.017
  54. 54. Rocca JD, Werner ME, Kramer SA, Huxford-Phillips RC, Sukumar R, Cummings ND, Vivero-Escoto JL, Wang AZ, Lin W. Polysilsesquioxane nanoparticles for triggered release of cisplatin and effective cancer chemoradiotherapy. Nanomedicine: Nanotechnology, Biology and Medicine. 2014;11:31-38. DOI: 10.1016/j.nano.2014.07.004
  55. 55. Maeda H, Bharate GY, Daruwalla J. Polymeric drugs for efficient tumor-targeted drug delivery based on EPR-effect. European Journal of Pharmaceutics and Biopharmaceutics. 2009;71:409-419. DOI: 10.1016/j.ejpb.2008.11.010
  56. 56. Sadhukha T, Prabha S. Encapsulation in nanoparticles improves anti-cancer efficacy of carboplatin. AAPS Pharmaceutical Science Technology. 2014;5:1029-1038. DOI: 10.1208/s12249-014-0139-2
  57. 57. Karanam V, Marslin G, Krishnamoorthy B, Chellan V, Siram K, Natarajan T, Bhaskar B, Franklin G. Poly(ε-caprolactone) nanoparticles of carboplatin: Preparation, characterization and in vitro cytotoxicity evaluation in U-87 MG cell lines. Colloids and Surfaces B: Biointerfaces. 2015;130:48-52. DOI: 10.1016/j.colsurfb.2015.04.005
  58. 58. Ahmeda F, Alib MJ, Kondapi AK. Carboplatin loaded protein nanoparticles exhibit improve anti-proliferative activity in retinoblastoma cells. International Journal of Biological Macromolecules. 2014;70:572-582. DOI: 10.1016/j.ijbiomac.2014.07.041
  59. 59. Khan MA, Zafaryab M, Mehdi SH, Quadri J, Rizvi MM. Characterization and carboplatin loaded chitosan nanoparticles for the chemotherapy against breast cancer in vitro studies. International Journal of Biological Macromolecules. 2017;97:115-122. DOI: 10.1016/j.ijbiomac.2016.12.090
  60. 60. Balas M, Constanda S, Duma-Voiculet A, Prodana M, Hermenean A, Pop S, Demetrescu I, Dinischiotu A. Fabrication and toxicity characterization of a hybrid material based on oxidized and aminated MWCNT loaded with carboplatin. Toxicology in Vitro. 2016;37:189-200. DOI: 10.1016/j.tiv.2016.09.011
  61. 61. Makharza S, Vittorio O, Cirillo G, Oswald S, Hinde E, Kavallaris M, Büchner B, Mertig M, Hampel S. Graphene oxide–Gelatin nanohybrids as functional tools for enhanced carboplatin activity in neuroblastoma cells. Pharmaceutical Research. 2015;32:2132-2143. DOI: 10.1007/s11095-014-1604-z
  62. 62. Yang C, Fu ZX. Liposomal delivery and polyethylene glycol‑liposomal oxaliplatin for the treatment of colorectal cancer. Biomedical Reports. 2014;2:335-339. DOI:10.3892/br.2014.249
  63. 63. Stathopoulos GP, Boulikas T, Kourvetaris A, Stathopoulos J. Liposomal oxaliplatin in the treatment of advanced cancer: A phase I study. Anticancer Research. 2006;26:1489-1493
  64. 64. Shi M, Fortin D, Paquette B, Sanche L. Convection-enhancement delivery of liposomal formulation of oxaliplatin shows less toxicity than oxaliplatin yet maintains a similar median survival time in F98 glioma-bearing rat model. Investigational New Drugs. 2016;34:269-276. DOI: 10.1007/s10637-016-0340-0
  65. 65. Suzuki R, Takizawa T, Kuwata Y, Mutoh M, Ishiguro N, Utoguchi N, Shinohara A, Eriguchi M, Yanagie H, Maruyama K. Effective anti-tumor activity of oxaliplatin encapsulated in transferrin–PEG–liposome. International Journal of Pharmaceutics. 2008;346:143-150. DOI: 10.1016/j.ijpharm.2007.06.010
  66. 66. Abu-Lila AS, Kiwada H, Ishida T. Selective delivery of oxaliplatin to tumor tissue by nanocarrier system enhances overall therapeutic efficacy of the encapsulated oxaliplatin. Biologicaland Pharmaceutical Bulletin. 2014:37;206-211. DOI:10.1248/bpb.b13-00540
  67. 67. Abu-Lila A, Suzuki T, Doi Y, Ishida T, Kiwada H. Oxaliplatin targeting to angiogenic vessels by PEGylated cationic liposomes suppresses the angiogenesis in a dorsal air sac mouse model. Journal of Controlled Release. 2009;134:18-25. DOI:10.1016/j.jconrel.2008.10.018
  68. 68. Oberoi HS, Nukolova NV, Kabanov AV, Bronich TK. Nanocarriers for delivery of platinum anticancer drugs. Advanced Drug Delivery Reviews. 2013;65:1667-1685. DOI: 10.1016/j.addr.2013.09.014
  69. 69. Wang R, Hu X, Xiao H, Xie Z, Huang Y, Jing X. Polymeric dinulcear platinum(II) complex micelles for enhanced antitumor activity. Journal of Materials Chemistry B. 2013;1:744-748. DOI: 10.1039/C2TB00240J
  70. 70. Cabral H, Nishiyama N, Kataoka K. Optimization of (1,2-diaminocyclohexane)platinum(II)-loaded polymeric micelles directed to improved tumor targeting and enhanced antitumor activity. Journal of Controlled Release. 2007;121:146-155. DOI:10.1016/j.jconrel.2007.05.024
  71. 71. Rafi M, Cabral H, Kano MR, Mi P, Iwata C, Yashiro M, Hirakawa K, Miyazono K, Nishiyama N, Kataoka K. Polymeric micelles incorporating (1,2-diaminocyclohexane)platinum(II) suppress the growth of orthotopicscirrhous gastric tumors and their lymph node metastasis. Journal of Controlled Release. 2012;159:189-196. DOI: 10.1016/j.jconrel.2012.01.038
  72. 72. Nowotnik DP, Cvitkovic E. ProLindac (AP5346): A review of the development of an HPMA DACH platinum polymer therapeutic. Advanced Drug Delivery Reviews. 2009;61:1214-1219. DOI:10.1016/j.addr.2009.06.004
  73. 73. Apps MG, Choi EH, Wheate NJ. The state of-play and future of platinum drugs. Endocrine-Related Cancer. 2015;22:R219–R233. DOI: 10.1530/ERC-15-0237
  74. 74. Song H, Xiao H, Zheng M, Qi R, Yan L, Jing X. A biodegradable polymer platform for co-delivery of clinically relevant oxaliplatin and gemcitabine. Journal of Materials Chemistry B. 2014;2:6560-6570. DOI: 10.1039/C4TB00678J
  75. 75. Xiao H, Li W, Qi R, Yan L, Wang R, Liu S, Zheng Y, Xie Z, Huang Y, Jing X. Co-delivery of daunomycin and oxaliplatin by biodegradable polymers for safer and more efficacious combination therapy. Journal of Controlled Release. 2012;163:304-314. DOI: 10.1016/j.jconrel.2012.06.004
  76. 76. Jain A, Jain SK, Ganesh N, Barve J, Beg AM. Design and development of ligand-appended polysaccharidic nanoparticles for the delivery of oxaliplatin in colorectal cancer. Nanomedicine. 2010;6:179-190. DOI: 10.1016/j.nano.2009.03.002
  77. 77. Tummala S, Gowthamarajan K, Satish Kumar MN, Wadhwani A. Oxaliplatin immuno hybrid nanoparticles for active targeting: An approach for enhanced apoptotic activity and drug delivery to colorectal tumors. Drug Delivery. 2016;23:1773-1787. DOI: 10.3109/10717544.2015.1084400
  78. 78. Vivek R, Thangam R, Nipunbabu V, Ponraj T, Kannan S. Oxaliplatin-chitosan nanoparticles induced intrinsic apoptoticsignaling pathway: A “smart” drug delivery system to breastcancer cell therapy. International Journal of Biological Macromolecules. 2014;65:289-297. DOI: 10.1016/j.ijbiomac.2014.01.054
  79. 79. Chen Y, Ding D, Mao Z, He Y, Hu Y, Wu W, Jiang X. Synthesis of hydroxypropylcellulose-poly(acrylic acid) particles with semi-interpenetrating polymer network structure. Biomacromolecules. 2008;9:2609-2614.DOI: 10.1021/bm800484e
  80. 80. Dutta RK, Sahu S. Development of oxaliplatin encapsulated in magnetic nanocarriers of pectin as a potential targeted drug delivery for cancer therapy. Results in Pharma Sciences. 2012;2:38-45. DOI: 10.1016/j.rinphs.2012.05.001
  81. 81. Wu L, Man C, Wang H, Lu X, Ma Q, Cai Y, Ma W. PEGylated multi-walled carbon nanotubes for encapsulation and sustained release of oxaliplatin. Pharmaceutical Research. 2013;30:412-423. DOI: 10.1007/s11095-012-0883-5
  82. 82. Lee PC, Lin CY, Peng CL, Shieh MJ. Development of a controlled-release drug delivery system by encapsulating oxaliplatin into SPIO/MWNT nanoparticles for effective colon cancer therapy and magnetic resonance imaging. Biomaterials Science. 2016;4:1742-1753. DOI:10.1039/c6bm00444j
  83. 83. Brown SD, Nativo P, Smith J-A, Stirling D, Edwards PR, Venugopal B, Flint DJ, Plumb JA, Graham D, Wheate NJ. Gold nanoparticles for the improved anticancer drug delivery of the active component of oxaliplatin. Journal of the American Chemical Society. 2010; 132:4678-4684. DOI: 10.1021/ja908117a
  84. 84. He H, Xiao H, Kuang H, Xie Z, Chen X, Jing X, Huang Y. Synthesis of mesoporous silica nanoparticle–oxaliplatin conjugates for improved anticancer drug delivery. Colloids and Surfaces B: Biointerfaces. 2014;117:75-81. DOI: 10.1016/j.colsurfb.2014.02.014
  85. 85. Della Rocca J, Huxford RC, Comstock-Duggan E, Lin W. Polysilsesquioxane nanoparticles for targeted platin-based cancer chemotherapy by triggered release. Angewandte Chemie, International Edition. 2011;50:10330-10334. DOI: 10.1002/anie.201104510
  86. 86. Bergamo A, Gagliardi R, Scarcia V, Furlani A, Alessio E, Mestroni G, Sava G. In vitro cell cycle arrest, in vivo action on solid metastasizing tumors, and host toxicity of the antimetastatic drug NAMI-A and cisplatin. Journal of Pharmacology and Experimental Therapeutics. 1999;289:559-564
  87. 87. Dragutan I, Dragutan V, Demonceau A. Editorial of special issue ruthenium complex: The expanding chemistry of the ruthenium complexes. Molecules. 2015;20:17244-17274. DOI: 10.3390/molecules200917244
  88. 88. Bergamo A, Gaiddon C, Schellens JHM, Beijnen JH, Sava G. Approaching tumour therapy beyond platinum drugs: Status of the art and perspectives of ruthenium drug candidates. Journal of Inorganic Biochemistry. 2012;106:90-99. DOI: 10.1016/j.jinorgbio.2011.09.030
  89. 89. Rademaker-Lakhai JM, Van Den Bongard D, Pluim D, Beijnen JH, Schellens JHM. A phase I and pharmacological study with imidazolium-trans-DMSO-imidazole-tetrachlororuthenate, a novel ruthenium anticancer agent. Clinical Cancer Research. 2004;10: 3717-3727. DOI: 10.1158/1078-0432.CCR-03-0746
  90. 90. Leijen S, Burgers SA, Baas P, Pluim D, Tibben M, van Werkhoven E, Alessio E, Sava G, Beijnen JH, Schellens JH. Phase I/II study with ruthenium compound NAMI-A and gemcitabine in patients with non-small cell lung cancer after first line therapy. Investigational New Drugs. 2015;33:201-214. DOI: 10.1007/s10637-014-0179-1
  91. 91. Hartinger CG, Zorbas-Seifried S, Jakupec MA, Kynast B, Zorbas H, Keppler BK. From bench to bedside–preclinical and early clinical development of the anticancer agent indazolium trans-[tetrachlorobis(1H-indazole)ruthenate(III)] (KP1019 or FFC14A). Journal of Inorganic Biochemistry. 2006;100:891-904. DOI: 10.1016/j.jinorgbio.2006.02.013
  92. 92. Hartinger CG, Jakupec MA, Zorbas-Seifried S, Groessl M, Egger A, Berger W, Zorbas H, Dyson PJ, Keppler BK. KP1019, a new redox-active anticancer agent-preclinical development and results of a clinical phase I study in tumor patients. Chemistry & Biodiversity. 2008;5:2140-2155. DOI: 10.1002/cbdv.200890195
  93. 93. Lentz F, Drescher A, Lindauer A, Henke M, Hilger RA, Hartinger CG, Scheulen ME, Dittrich C, Keppler BK, Jaehde U. Pharmacokinetics of a novel anticancer ruthenium complex (KP1019, FFC14A) in a phase I dose-escalation study. Anti-Cancer Drugs. 2009;20:97-103. DOI: 10.1097/CAD.0b013e328322fbc5
  94. 94. Mestroni G, Alessio E, Sava G, Pacor S, Coluccia M, Boccarelli A. Water-soluble ruthenium(III)-dimethyl sulfoxide complexes: Chemical behaviour and pharmaceutical properties Met. Based Drugs. 1994;1:41-63. DOI: 10.1155/MBD.1994.41
  95. 95. Ott I, Gust R. Non platinum metal complexes as anti-cancer drugs. Archiv der Pharmazie. 2007;340:117-126. DOI: 10.1002/ardp.200600151
  96. 96. Levina A, Aitken JB, Gwee YY, Lim ZJ, Liu M, Singharay AM, Wong PF, Lay PA. Biotransformations of anticancer ruthenium(III) complexes: An X-ray absorption spectroscopic study. Chemistry--A European Journal. 2013;19:3609-3619. DOI: 10.1002/chem.201203127
  97. 97. Blunden BM, Stenzel MH. Incorporating ruthenium into advanced drug delivery carriers – An innovative generation of chemotherapeutics. Journal of Chemical Technology and Biotechnology. 2015;90: 1177-1195. DOI: 10.1002/jctb.4507
  98. 98. Fischer B, Heffeter P, Kryeziu K, Gille L, Meier SM, Berger W, Kowol CR, Keppler BK. Poly(lactic acid) nanoparticles of the lead anticancer ruthenium compound KP1019 and its surfactant-mediated activation. Dalton Transactions. 2014;43:1096-1104. DOI: 10.1039/c3dt52388h.
  99. 99. He L, Huang Y, Zhu H, Pang G, Zheng W, Wong Y-S, Chen T. Cancer-targeted monodisperse mesoporous silica nanoparticles as Carrier of ruthenium polypyridyl complexes to enhance theranostic effects. Advanced Functional Materials. 2014;24:2754-2763. DOI: 10.1002/adfm.201303533
  100. 100. Rojas S, Quartapelle-Procopio E, Carmona FJ, Romero MA, Navarro JAR, Barea E. Biophysical characterisation, antitumor activity and MOF encapsulation of a half-sandwich ruthenium(II) mitoxantronato system. Journal of Materials Chemistry B. 2014;2:2473-2477. DOI: 10.1039/C3TB21455A
  101. 101. Blunden BM, Rawal A, Lu H, Stenzel MH. Superior chemotherapeutic benefits from the ruthenium-based anti-metastatic drug NAMI-A through conjugation to polymeric micelles. Macromolecules. 2014;47:1646-1655. DOI: 10.1021/ma402078d
  102. 102. Vaccaro M, Del Litto R, Mangiapia G, Carnerup AM, D'Errico G, Ruffo F, Paduano L. Lipid based nanovectors containing ruthenium complexes: A potential route in cancer therapy. Chemical Communications. 2009;1404-1406. DOI: 10.1039/B820368G
  103. 103. Mangiapia G, D'Errico G, Simeone L, Irace C, Radulescu A, Di Pascale A, Colonna A, Montesarchio D, Paduano L. Ruthenium-based complex nanocarriers for cancer therapy. Biomaterials. 2012;33:3770-3782. DOI: 10.1016/j.biomaterials.2012.01.057
  104. 104. Simeone L, Mangiapia G, Vitiello G, Irace C, Colonna A, Ortona O, Montesarchio D, Paduano L. Cholesterol-based nucleolipid-ruthenium complex stabilized by lipid aggregates for antineoplastic therapy. Bioconjugate Chemistry. 2012;23:758-770. DOI: 10.1021/bc200565v
  105. 105. Montesarchio D, Mangiapia G, Vitiello G, Musumeci D, Irace C, Santamaria R, D’Erricoa G, Paduano L. A new design for nucleolipid-based Ru(III) complexes as anticancer agents. Dalton Transactions. 2013;42:16697-16708. DOI: 10.1039/c3dt52320a.
  106. 106. Mangiapia G, Vitiello G, Irace C, Santamaria R, Colonna A, Angelico R, Radulescu A, D'Errico G, Montesarchio D, Paduano L. Anticancer cationic ruthenium nanovectors: From rational molecular design to cellular uptake and bioactivity. Biomacromolecules. 2013;14:2549-2560. DOI: 10.1021/bm400104b
  107. 107. Vitiello G, Luchini A, D'Errico G, Santamaria R, Capuozzo A, Irace C, Montesarchio D, Paduano L. Cationic liposomes as efficient nanocarriers for the drug delivery of an anticancer cholesterol-based ruthenium complex. Journal of Materials Chemistry B. 2015;3:3011-3023. DOI: 10.1039/C4TB01807A
  108. 108. Riccardi C, Musumeci D, Irace C, Paduano L, Montesarchio D. RuIII complexes for anticancer therapy: The importance of being nucleolipidic. European Journal of Organic Chemistry. 2016. DOI: 10.1002/ejoc.201600943
  109. 109. Servin P, Laurent R, Gonsalvi L, Tristany M, Peruzzini M, Majoral J-P, Caminade A-M. Grafting of water-soluble phosphines to dendrimers and their use in catalysis: positive dendritic effects in aqueous media. Dalton Transactions. 2009;38:4432-4434. DOI: 10.1039/B906393P
  110. 110. Valente A, Garcia MH. Syntheses of macromolecular ruthenium compounds: A new approach for the search of anticancer drugs. Inorganics. 2014;2:96-114; DOI: 10.3390/inorganics2010096
  111. 111. Govender P, Therrien B, Smith GS. Bio-metallodendrimers–Emerging strategies in metal-based drug. European Journal of Inorganic Chemistry. 2012;2012:2853-2862. DOI: 10.1002/ejic.20120016

Written By

Valentina Uivarosi, Rodica Olar and Mihaela Badea

Submitted: 21 October 2016 Reviewed: 03 March 2017 Published: 23 August 2017