Next Article in Journal
Erbium-Doped Lu2O3-MgO and Sc2O3-MgO IR-Transparent Composite Ceramics
Next Article in Special Issue
Tin/Tin Oxide Nanostructures: Formation, Application, and Atomic and Electronic Structure Peculiarities
Previous Article in Journal
Effect of Biosynthesized Silver Nanoparticles on the Growth of the Green Microalga Haematococcus pluvialis and Astaxanthin Synthesis
Previous Article in Special Issue
Nanoparticle and Nanostructure Synthesis and Controlled Growth Methods
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Platinum-Based Nanoformulations for Glioblastoma Treatment: The Resurgence of Platinum Drugs?

by
Paula Alfonso-Triguero
1,2,
Julia Lorenzo
1,3,
Ana Paula Candiota
1,3,4,
Carles Arús
1,3,4,
Daniel Ruiz-Molina
2 and
Fernando Novio
2,5,*
1
Institut de Biotecnologia i de Biomedicina, Departament de Bioquimica i Biologia Molecular, Universitat Autònoma de Barcelona, 08193 Cerdanyola del Vallès, Spain
2
Catalan Institute of Nanoscience and Nanotechnology (ICN2), CSIC and BIST, Campus UAB, Bellaterra, 08193 Barcelona, Spain
3
Departament de Bioquímica i Biologia Molecular, Universitat Autònoma de Barcelona, 08193 Cerdanyola del Vallès, Spain
4
Centro de Investigación Biomédica en Red, Bioingeniería, Biomateriales y Nanomedicina (CIBER-BBN), 08193 Cerdanyola del Vallès, Spain
5
Departament de Química, Universitat Autònoma de Barcelona (UAB), Campus UAB, 08193 Cerdanyola del Vallès, Spain
*
Author to whom correspondence should be addressed.
Nanomaterials 2023, 13(10), 1619; https://0-doi-org.brum.beds.ac.uk/10.3390/nano13101619
Submission received: 12 April 2023 / Revised: 6 May 2023 / Accepted: 9 May 2023 / Published: 12 May 2023
(This article belongs to the Special Issue Current Review in Nanofabrication and Nanomanufacturing)

Abstract

:
Current therapies for treating Glioblastoma (GB), and brain tumours in general, are inefficient and represent numerous challenges. In addition to surgical resection, chemotherapy and radiotherapy are presently used as standards of care. However, treated patients still face a dismal prognosis with a median survival below 15–18 months. Temozolomide (TMZ) is the main chemotherapeutic agent administered; however, intrinsic or acquired resistance to TMZ contributes to the limited efficacy of this drug. To circumvent the current drawbacks in GB treatment, a large number of classical and non-classical platinum complexes have been prepared and tested for anticancer activity, especially platinum (IV)-based prodrugs. Platinum complexes, used as alkylating agents in the anticancer chemotherapy of some malignancies, are though often associated with severe systemic toxicity (i.e., neurotoxicity), especially after long-term treatments. The objective of the current developments is to produce novel nanoformulations with improved lipophilicity and passive diffusion, promoting intracellular accumulation, while reducing toxicity and optimizing the concomitant treatment of chemo-/radiotherapy. Moreover, the blood–brain barrier (BBB) prevents the access of the drugs to the brain and accumulation in tumour cells, so it represents a key challenge for GB management. The development of novel nanomedicines with the ability to (i) encapsulate Pt-based drugs and pro-drugs, (ii) cross the BBB, and (iii) specifically target cancer cells represents a promising approach to increase the therapeutic effect of the anticancer drugs and reduce undesired side effects. In this review, a critical discussion is presented concerning different families of nanoparticles able to encapsulate platinum anticancer drugs and their application for GB treatment, emphasizing their potential for increasing the effectiveness of platinum-based drugs.

1. Introduction

Glioma is a general term used to describe primary brain tumours with an imbalance between high mortality and morbidity compared to their low incidence. These tumours are classified according to their originating cell (supportive glial cells) including astrocytomas, oligodendrogliomas, ependymomas, and mixed gliomas. Glioblastomas (GB) are the most malignant and frequent type of primary astrocytomas, accounting for about 33% of all brain tumours and about 80% of the total malignant central nervous system (CNS) tumours in adults [1]. Despite several efforts to improve GB treatment, nowadays it is still a deadly disease with poor prognosis and a median survival of less than 2 years upon diagnosis and a 5-year survival rate of 5.1% [2]. Moreover, GB is characterized by a diffuse infiltration of the adjacent brain parenchyma and development of drug resistance to standard treatments [3].
In addition to GB tumour cells, it is important to consider the tumour microenvironment (TME) factors in which neoplastic and non-neoplastic (e.g., tumour-associated macrophages, infiltrating lymphocytes) cell types interact, influencing GB growth, progression, and therapy resistance [4,5]. Recent studies indicate that differentiated tumour cells may dedifferentiate, acquiring a stem-like phenotype, in response to microenvironment stressful conditions such as hypoxia, acidic extracellular pH, or presence of nitric oxide [6]. Moreover, patient habits such as smoking also can induce pharmacokinetic interactions, through prevention of proapoptotic events or downregulation of proapoptotic proteins, and have a great impact on the effectiveness and toxicity of anticancer drugs [7]. As a result, adjustments of drug dosages should be taken into account in specific cases.
The standard of care for many years has consisted of surgical resection followed by radiotherapy (RT) and concomitant chemotherapy (CT), first described by Stupp et al. [8]. However, due to the location and infiltrative nature of GB, complete surgical resection of the tumour is often not feasible without a high risk of neurological damage for the patient [9]. The main anticancer drug used for GB treatment is temozolomide (TMZ), although acquired resistance of TMZ often contributes to the poor efficacy of the treatment, and patient survival is only slightly improved in few months. After almost two decades since the initial publication, and several efforts devoted in preclinical and clinical studies, the harsh truth is that the outcome of GB therapeutic approaches has not significantly improved.
Thus, the failure in GB treatment by now is attributable to different challenges during therapy, and also to interactions among different cells within the tumour microenvironment (TME). Therefore, methods to monitor genomic heterogeneity and even immune evasion would be interesting to adequate the therapy, using non-invasive strategies, to improve delivery across the blood–brain barrier and facilitate the success of personalized therapies [10]. On top of that, the economic burden of GB treatment has been reviewed in terms of cost-of-illness and cost-effectiveness studies [11].

2. Current Landscape for GB Treatment

The most effective standard clinical protocol for GB used nowadays was established back in 2005 [8] and, as previously stated, involves maximum surgical resection without compromising neurological functions followed by a combination of CT (oral administration of TMZ) and RT, and a period of adjuvant TMZ. However, there are several challenges related to treatment effectiveness: the infiltrative GB nature hampers complete tumour removal, since tumour excessive resection can lead to brain dysfunction, while incomplete removal may lead to tumour regrowth from remaining tumour cells. Moreover, a key point to brain tumour treatment is the contribution of the biological barriers, blood–brain barrier (BBB) and blood–tumour barrier (BTB) to restrictions in local drug delivery.
TMZ, approved by the FDA in 1999, is the most commonly used anticancer drug for GB treatment due to its lipophilic nature enabling it to easily cross the BBB [12]. Although TMZ demonstrated increasing patient survival and suppressing tumour growth in early stages of disease, ca. 50% of treated patients do not respond to this anticancer drug. Additionally, several GB cell lines such as U87, U251, U373, or T98G have been shown to develop TMZ resistance [13,14,15]. The combination of TMZ with radiotherapy does not significantly improve GB patient survival [16]. Then, even after aggressive therapy, overall survival is below 2 years and therapeutic alternatives are rather limited since surgery is not always feasible [17]. Approaches such as tumour-treating fields (TTFields), in which alternating electrical fields are delivered via cutaneous transducer arrays, inhibited GB cell proliferation by interfering with mitotic apparatus, although a recent review highlighted that other effects related to TTFields such as immunogenic signaling or anti-migratory effects can be an added value in GB treatment [18].
Alternative chemotherapeutic approaches offered as second-line treatment upon TMZ failure include nitrosoureas such as lomustine, carmustine, fotemustine [19], and antiangiogenic agents which are human monoclonal antibodies that inhibits vascular endothelial growth factor (VEGF), such as Bevacizumab [20]. Additionally, many research efforts have been described towards targeted therapies in GB: targeting epidermal growth factor receptor (EGFR) [21], inhibiting protein kinase C (PKC) [22], or inhibiting phosphatidylinositol 3-kinase (PI3K) and mammalian target of rapamycin (mTOR) signaling pathways PI3K and mTor [23], respectively, among others. Considering the relevant immunosuppressive effects on the GB microenvironment and crosstalk with TME, immunotherapeutic approaches are also being currently exploited, alone or in combination with standard/novel GB therapies. These therapeutic actions have been basically focused on immune checkpoint inhibitors such as anti-PD-1 and anti-CTLA-4 or dendritic cell-based vaccine therapies, as well as adoptive T-cell therapies [24]. Interestingly, Liu et al. reported the co-encapsulation of TMZ and an immunity activator (OTX015), using ApoE decorated red blood cell membrane, to obtain a biomimetic nanomedicine. The resulting nanoformulation showed a brain-targeted drug co-delivery and synergistic chemoimmunotherapy in vivo, with marked tumour inhibition and enhanced anti-tumour immune responses [25]. However, the strong immunosuppressive environment in GB may require further additional strategies to sensitize GB cells to immunotherapy [26]. Altogether, these factors contribute to the lack of a cure for GB, requiring continuous research efforts towards novel treatment strategies.
An additional key factor to take into account is the side effects implication of the chemotherapeutics, and the interaction with concurrent medications. Although TMZ is one of the main chemotherapeutics used for glioblastoma first-line treatment, the administration of intense doses used in the clinical protocol provoke relevant side effects such as bone marrow problems (leading to the need of treatment halting due to anemia, thrombocytopenia, etc.) [27]. On the other hand, side effects of nitrosoureas are related to bone marrow, as well as liver, kidney, and lung damage, and an increased risk of developing leukemia [28]. In addition, nitrosoureas can cross the blood–brain barrier and cause neurotoxicity. Platinum complexes, and more specifically cisplatin-related therapeutics, are usually associated with nephrotoxicity and neurotoxicity [29]. Any type of therapy directed towards fast-proliferating cells might affect cell populations such as digestive system, skin, and immune cells, among others. Some therapeutic combinations have demonstrated to be effective and act synergistically for GB treatment. Moreover, since combination drugs target multiple pathways, their administration can cause significant savings: smaller drug doses, lower treatment failure rate, and slower development of drug resistance. However, it is important to consider some limitations to the use combination therapy in GB since it can trigger cumulative side effects and modest clinical benefit [30]. In this scenario, it is important to build mathematical models of synergism/antagonism of drugs and pathways for the prediction of drug combinations for the treatment of heterogeneous tumours in GB. It is also important to identifying synergistic interactions between chemotherapy, radiotherapy, and immunotherapy in order to maximize the antitumour potential of individual treatment approaches.

3. Platinum and Glioblastoma: State-of-the-Art

Platinum (Pt) complexes arise as an alternative option for treating brain tumours through chemotherapy. Platinum complexes are alkylating-like drugs that crosslink with the DNA, interfering with its repairing mechanism and inducing DNA damage and cell apoptosis [31]. Pt derivatives may also modulate anti-tumour immunity [32], impair tumour invasiveness through matrix metalloproteinase (MMP) downregulation, exhibit anti-angiogenic effects, and modulate MGMT DNA repair enzyme, among others. Pt-based drugs and prodrugs have demonstrated to be effective as anticancer agents in a variety of carcinomas (i.e., testicular, ovarian, lung, and head and neck cancer) [33]. The gold-standard Pt-based anticancer drugs cisplatin, carboplatin, and oxaliplatin have been proved to exert clinical effects in brain tumour patients [29,34,35,36]. Three additional platinum complexes (nedaplatin, lobaplatin, and heptaplatin) have been approved in specific countries. Figure 1 depicts the chemical structures of these approved Pt drugs [37]. This section may be divided by subheadings. It should provide a concise and precise description of the experimental results, their interpretation, as well as the experimental conclusions that can be drawn.
Pt complexes are described as effective therapeutic agents against gliomas [38,39,40] showing the following: (I) anti-angiogenic effects [41], (II) enhanced adjuvant therapy efficacy (TMZ and radiotherapy) [36,42,43], (III) successful combination with a tyrosine kinase inhibitor [44], and (IV) acceptable toxicity in chemotherapy-naïve patients with recurrent GB [45]. Moreover, the well know chemistry of Pt allows designing and generating an enormous variety of platinum-based complexes with stimuli-responsiveness properties in front of pH variation, redox activity, temperature changes, light irradiation, or enzyme overexpression [46]. These developments offer a good opportunity for obtaining site-specific prodrugs to maximize the therapeutic efficacy and minimize the side effect of platinum metallodrugs in anticancer therapies. However, reports in the literature describing the use of Pt, either alone or in combination in different approaches/scenarios, have controversial results. In vitro studies with human SNB19 and U87 GB cells reported the use of non-conventional Pt complexes (platinum-acridine hybrid agent) with excellent results and toxicity profiles superior to cisplatin [47]. Different therapeutic approaches have been tested—mostly in vitro—in combination with Pt such as bee venom [48], β-elemene extracted from curcuma wenyujin [49], nitrosoureas [50,51], bioconjugates combining EGFR targeting [52], microRNAs (miRNAs) [53,54], protein kinases C [55], tumour necrosis factor (TNF) [56], a histone deacetylase inhibitor (HDACi) [57], topoisomerase II [58], PI3K inhibitors [59], and a BK (large conductance, Ca2+-activated K+) channel inhibitor [60]. In addition, a combination of p(65)+Be neutrons irradiation plus cisplatin was attempted in U87 GB cells, resulting in a marked reinforcement of cytotoxic effect [61].
Regarding in vivo approaches, both in preclinical and clinical settings, single-agent carboplatin/cisplatin has been attempted as a ‘rescue’ treatment for high-grade glioma (HGG)-afflicted patients who did not respond after treatment with chemotherapy, nitrosoureas, or temozolomide, showing only discrete, non-significant improvement in outcome [62]. When coming to drug combination, a phase II trial combining TMZ and cisplatin was performed in pretreated and recurring HGG patients (GB and grade III gliomas treated with standard surgery + chemo-radiotherapy) [35]. This study reported a progression-free survival of 35% (at 6 months) and 13.8% (at 12 months), although grades 4 and 5 side effects were observed. Although results in this study were promising, the same authors recognized that they originated from a small, non-randomized cohort, and a more detailed studied would be needed to achieve confident conclusions in this regard. Pt complexes have also been explored in combination with other chemotherapy approaches, such as a dose-intense TMZ [63,64], both in preclinical models [65] and phase II trials [66]. Moreover, Rousseau et al. reported the combination of cisplatin (delivered via convection-enhanced delivery (CED)) with photon irradiation leading to enhanced survival of F98 glioma-bearing rats [67]. Thus, Pt drugs may have great therapeutic potential when effectively delivered to the tumour regions [68]. Despite the favorable results previously described, in practice, Pt therapy is considered fourth-line chemotherapeutics, i.e., it will only be used when all standard therapeutic protocols fail. At this point, we are probably facing unfavorable scenarios, dealing with highly resistant tumours. The performance observed in clinical settings has limited incidence on the overall survival of patients with brain tumours [69], neither alone nor in combination with radiotherapy [70,71] or chemotherapeutics such as carmustine [71] and TMZ [72].
Despite initial benefits reported by some authors, the administration of such Pt(II) complexes is often associated with severe systemic toxicity resulting from long-term treatment [73]. Moreover, the BBB contributes to the scarce drug arrival to the brain when drugs are administered orally or intravenously (i.v.). This is a key factor to explain the poor activity of such complexes in brain tumours, since passive diffusion is uncommon, and it is only feasible for small lipophilic compounds or endogenous molecules that pass through specific transporters. In fact, the use of receptor-mediated transport methodologies has been studied for many years as the most efficient mechanism for drug delivery to the brain [74]. Different methods have been explored to minimize the BBB-related challenges and to overcome this limited uptake of drugs in brain tumour while minimizing side effects. These strategies include approaches to increase the BBB permeability [75,76], use of biodegradable implants directly in the tumour [77,78], or more aggressive approaches, such as the use of a catheter to deliver drugs directly into the brain through CED [79].
Despite successful results, the development of formulations that need local device placements for drug delivery are not attractive from a clinical point of view, so new approaches ensuring proper and efficient chemotherapeutic delivery still represent a major challenge nowadays [37]. The use of nanoparticles (NPs) to increase the i.v. local delivery of drugs into the brain is one of the most promising approaches [80,81]. Their small size and large surface area are suitable to increase the solubility and bioavailability of the anticancer drugs, facilitating BBB diffusion and concentrating higher CT doses in the brain parenchyma [82]. Although NPs can benefit from the enhanced permeability and retention (EPR) effect to access and be retained in tumour tissues, the possibility of their surface functionalization enables targeting receptors or specific (bio)molecules is an added value to increase specificity or to evade the mononuclear phagocyte system [83,84,85]. Additionally, NPs can transport different therapeutic agents at the same time, protecting them from premature metabolism/degradation and inducing a precise control of the drug release [86]. All these advantages convert nanoparticles into suitable systems to maximize therapeutic efficacy of anticancer drugs while reducing their undesirable toxic effects [87].
Specifically, for minimizing drug resistance and side effects of platinum-based anticancer drugs/prodrugs while increasing their efficacy, efficient delivery systems based on cancer-specific targeting can be used [88,89]. Thus, different drug delivery systems including liposomes [90], dendrimers [52,91], polymers [92], nanotubes [93], or inorganic [94,95] and hybrid [96] nanoparticles have been evaluated with these purposes. Among them, liposomes are one of the most developed and promising drug carriers for platinum-based drugs [97], showing excellent effects without toxicity in phase I clinical trials [98]. It has been demonstrated that the nanoformulation of platinum-based drugs can significantly reduce the drug toxicity [99], improving drug delivery to tumours with a concomitant reduction in glioma growth, without neurotoxicity, and showing increased survival rate [100]. Moreover, the combination of drug administration using nanoformulations with other methodologies to improve brain delivery such as the disruption of the BBB by non-invasive transient focused ultrasound methods has been described [101,102]. Due to the complexity and heterogeneity of tumours [103], it is suggested that each patient might need personalized and combined treatments to increase the efficacy of platinum-based anticancer drugs against brain tumours.

4. Platinum-Based Drugs Limitations and Opportunities

Many efforts have been made to improve brain tumour treatment efficacy. The clinical treatment using platinum-based anticancer drugs reveals a series of limitations related to their clinical application and to the signaling mechanisms that cause drug resistance [29]. However, different studies have demonstrated that the therapeutic efficacy of platinum-based anticancer drugs can be significantly higher than TMZ efficacy. Thus, some questions arise, namely (i) what are the specific limitations related to the use of cisplatin in the clinical/preclinical practice of brain cancer treatment, and (ii) can these limitations be minimized? Below we will comment on some of them.

4.1. Factors Related to Brain Uptake

Less than 5% of drug plasma concentration is detected in the brain after i.v. delivery of platinum-based drugs to nonhuman primates, due to high selectivity of the BBB [36]. Even in the case of malignant brain tumours with compromised BBB integrity, only modest platinum compounds uptake was reported [104]. On the other hand, it has been demonstrated that chronic treatment with platinum-based drugs results in platinum accumulation in the brain. Different studies reveal that most of the adverse effects of chemotherapy are cumulative and occur after chronic exposure to the drug [105]. Along with cumulative chronic side effects, platinum drug treatment (i.e., oxaliplatin) also causes transient acute neuropathy, affecting sensory nerves in particular [106].

4.2. Narrow Therapeutic Window

Cisplatin is a potent chemotherapeutic agent with good results for standard medulloblastoma treatment but not for GB therapeutic protocols [107,108]. However, there is no clear explanation for the differences observed in the clinical efficacy against medulloblastomas and GB, even though cisplatin is effective in vitro against the latter. Recent results demonstrate that treatment with intratumoural cisplatin may be an efficient approach for brain tumour treatment with an effective narrow therapeutic window [109].

4.3. Inappropriate Dosing Schedule

Frequency of administration may be a key issue when balancing effects against tumour cells and effects over host immune system. Platinum-based chemotherapy can enhance host antitumour immune responses in several ways, e.g., inducing immunogenic cell damage (ICD), increasing the activity of tumour-killing immune cells, enhancing the sensitivity of tumour cells to immune checkpoint inhibitors [110], and increasing calreticulin and major histocompatibility complex (MHC I) expression in vivo [111]. With this in mind, unsuitable administration schedules, in addition to tumour killing, may also adversely affect the host immune system, leading to opposed effects. It has been shown that the immune-related effects of these chemotherapeutics can be dependent on the drug specificity, distribution, dosing, tumour model, and type of immunotherapy in case of drug combination [109].

4.4. Extensive Off-Target Toxicity

The use of platinum drugs for the treatment of GB has shown minimal success, in part due to the extensive off-target toxicities such as nephrotoxicity or neurotoxicity [112,113,114]. This is not an isolated event, considering that all large molecules and most of the small ones fail to reach the brain within the required therapeutic levels [115], even in GB showing disrupted BBB. Thus, achieving suitable therapeutic doses within brain tumours sometimes requires the use of adjuvants [116,117], which may also contribute to severe adverse effects. Strategies considered for increasing BBB permeability [118] include cisplatin delivery within biodegradable polymer implants into the tumour bed of patients [119] or bypassing the BBB via injection under CED [120]. However, since some of the aforementioned approaches are not feasible in clinical practice, the real improvement of half-life and toxicity is rather limited. Several studies indicate that free Pt-drugs at high doses induce lymphodepletion and may hinder rather than stimulate antitumour immune responses [121]. However, new emerging information points to the broad and multi-faceted therapeutic potential of platinum-based agents, improving the therapeutic ratio. Recent recognized immunomodulatory properties of platinum compounds seem to be able to overcome many of the mechanisms related to GB immune evasion [122]. The main advantages of nanoformulations are the site-specific accumulation in the brain tumour, minimizing the systemic toxicity from therapeutic drugs and reducing the off-target effects [123].

4.5. Sequestering/Deactivating Reactions

One of the main objectives in platinum-based therapies is to minimize unwanted side reactions with biomolecules prior to DNA binding. For this purpose, the development of Pt(IV) prodrugs (most of them obtained by oxidation of the Pt(II)-related complex) afforded complexes with fine-tune desired biological properties such as lipophilicity, redox stability, cancer-cell targeting, improved cellular uptake, reduced off-target toxicity, and in some cases capability to modify the mechanism of action of the Pt(II) counterpart [37]. Reduction in the inner Pt(IV) center to anticancer-active Pt(II) due to the intracellular reductive environment, in concert with the loss of two ligands, is thought to be essential for the anticancer activity of these agents. Moreover, the possibility of attaching additional ligands to the octahedral coordinative sphere of the Pt(IV) metal center allows modifying its physicochemical properties and also facilitates attachment to nanoparticles and other carrier systems. The use of Pt(IV) complexes also offers solutions to different deactivation/sequestration pathways that can reduce the therapeutic actuation of platinum complexes, preventing cancer cells from triggering apoptosis and inducing the platinum complexes’ resistance (see Figure 2) [124].
Apart from opportunities offered by the nanoformulations, the unique properties of drug delivery vehicles can additionally provide capabilities of in vivo tracking of encapsulated drugs [125]. Nanoparticles have demonstrated potential as dual imaging contrast agents for magnetic resonance imaging (MRI) and computed tomography (CT) in glioma diagnosis [126]. Considering the disappointing therapeutic landscape for GB patients, discovering novel and safe methods to encapsulate and enhance drug delivery overpassing the BBB while improving biodistribution/chemical stability and decreasing side effects, has overall become a medical priority.

5. Platinum-Based Drugs Limitations and Opportunities

The studies published to date concerning the use of platinum-based nanoformulations for GB treatment were grouped into different categories according to the type of drug used, as follows. A summary of the principal examples is shown in Table 1.

5.1. Cisplatin

Duan et al. summarized several cisplatin-based nanoformulations with potential for clinic translation [127]. Reports related to the use of NPs for brain tumour treatment date back at least 10 years, with the encapsulation of cisplatin in polymeric NPs [128,129,130,131], polymeric micelles [132,133,134], polymeric conjugates [135,136], dendrimers [137,138], liposomes [139,140,141], nanocapsules [142,143], or its integration into metallic nanoparticles [144,145,146,147], silica nanoparticles [148,149,150], or hybrid nanoparticles (i.e., carbon nanotubes, nanoscale coordination polymers) [96,151,152,153,154,155,156,157,158]. Most of the published examples showed improved antitumoural effects, enhanced BBB crossing, and decreased side effects in comparison with free drugs. These benefits are related to the intrinsic properties of the NPs that can avoid the drug systemic toxicity, modulate the release of the therapeutic agent, and increase its accumulation into the tumour area by passive or active targeting processes.
Different metal-based NPs containing cisplatin drugs designed for GB treatment were reported. Makharza et al. described a nanocarrier combining γ-Fe2O3 NPs with nanographene oxide (NGO) for the selective vectorization of cisplatin. While NGO conferred high loading capabilities for cisplatin, the magnetic properties of the nanoparticles provided their magnetic guidance for targeting and delivery of therapeutics. The combination resulted in a sustained in vitro drug release with therapeutic potential against human U87 GB cells with negligible toxicity, together with the possibility to spatially control the drug delivery in the site of action [159].
A couple of publications were reported in 2018 using gold nanoparticles (Au-NPs) for GB combined with radiation treatment. Coluccia et al. described that cisplatin conjugated Au-NPs chemotherapy was synergistic with radiation and MR-guided Focused Ultrasound (MRgFUS). Namely, viability assays with GB cell lines in vitro (U87, U251, T98G, U138) demonstrated cell growth inhibition compared to free cisplatin and showed marked synergy with radiation therapy, while in vivo studies showed increased BBB permeability and brain drug delivery through MRgFUS [160]. In another example, Gotov et al. also presented cisplatin conjugated Au-NPs coated with hyaluronic acid [161]. Similarly, NPs showed enhanced cytotoxicity activity in comparison with free cisplatin in human breast adenocarcinoma MCF-7 cells, human primary U87 GB cells, or murine fibroblast NIH/3T3 (control) cell lines (see Figure 3). In vivo antitumour efficacy was demonstrated in tumour models upon i.v. administration and subsequent exposure to a near infra-red laser in the tumour site. These drug-loaded NPs showed marked in vivo activity due to their (i) long time of circulation in blood stream and (ii) selective accumulation in the tumour. This formulation demonstrated an enhanced therapeutic effect and reduced toxicity of cisplatin combining chemotherapy and laser treatment.
In vivo experiments reported back in 2010 with the commercial agent Lipoplatin™, which had already reached phase II/III clinical studies for non-small-cell lung cancer (NSCLC) [162], HER2/neu negative metastatic breast tumour [163], and advanced gastric tumour, were not successful [164]. Unexpectedly, its administration using CED showed marked neurotoxicity resulting in death within a few days, while the i.v. administration was well tolerated. The same authors have also tested a home-made nanoformulation combining cisplatin with lipid cholesteryl hemisuccinate (CHEMS), with a Pt loading efficiency of 25%. After 24 h treatment, CHEMS showed higher in vitro cytotoxicity against F98 glioma cells in comparison with free cisplatin and an excellent intracerebral retention in F98 glioma-bearing rats. Unfortunately, 10–14 days after administration, CHEMS unveiled dose-dependent neuropathologic findings [165]. Nowadays, there is an intense effort to promote targeting upon NP surface functionalization with peptides, antibodies, or small molecules recognizing transporters, antigens, or receptors characteristic of tumoural cells [166]. Shein et al. described liposomes with sustained release of cisplatin reaching relevant intracellular concentration in U87 GB cells. This promising result was achieved thanks to the conjugation of liposomes with antibodies against the vascular endothelial growth factor (VEGF) and its receptor type II (VEGFR2), which account for resistance and rapid progression of brain tumours [167]. Later on, Ashrafzadeh et al. described pegylated liposomal formulations encapsulating cisplatin and decorated with thiolated OX26 monoclonal antibodies to target transferrin receptors (TR). These liposomes caused an increase in the cellular uptake by 1.4-fold in comparison to non-targeted analogous and an increase by 1.7-fold in the mean survival of the C6 glioma-bearing rats compared to non-targeted ones with notable reduction in toxicity effects [168].
Together with liposomes, polymeric NPs have been one of the most developed nanoformulations for cisplatin delivery in GB treatment. One of the first developments reported consisted in Pt-bearing Poly(Lactide-Co-Glycolide) (PLGA) NPs coated with protamine. These nanosystems were reported to cross the BBB using an in vitro model of bovine brain microvessel endothelial cell assay performed in 2014 [169]. In addition, the same study reported therapeutic activity against U87 human GB cells with these NPs. Similar poly(lactic-co-glycolic acid)-block-polyethyleneglycol NPs were used to target mitochondrial DNA in cisplatin-resistant cells. This was achieved due to their high encapsulation payloads and mitochondria-targeting abilities upon surface functionalization with a triphenylphosphonium cation [170]. More successful experiments were performed with biodegradable PLGA NPs in 2014, which provided cisplatin-controlled release, with brain penetration and subsequent tumour targeting observed in ex vivo human and murine brain samples [171]. Experiments reported by Shahmabadi et al. using cisplatin-loaded (25% encapsulation efficiency) polybutylcyanoacrylate (PBCA) nanoparticles layered with polysorbate 80 failed to cross the BBB, most likely due its diameter average—larger than 400 nm [172]. PEGylated-Poly(aspartic acid) NPs encapsulating cisplatin with deep tumour penetration after local administration by CED were described by Zhang et al. [173]. With these brain-penetrating NPs, cisplatin reached concentrations feasible to kill tumour cells without healthy brain toxicity. This was demonstrated in in vivo experiments with rats bearing F98 orthotopic gliomas whose median survival was significantly increased in comparison to animals treated with free cisplatin. Later on, in 2018, poly(ethylene oxide)-triblock polymers were combined with Gd3+/cisplatin to obtain micelles which can act as contrast agents in MRI acquisitions while increasing the therapeutic effect of free cisplatin. In fact, the formulated prodrug exhibited up to 50-fold increased accumulation in human GB cell lines and up to 32-fold enhanced subsequent Pt-DNA adduct formation in comparison with free cisplatin [174]. Moreover, in vivo MRI monitoring of Gd-bearing nanoparticles within the brain after CED determined their promising potential as multifunctional drug delivery systems for both therapy and MRI tracking. Other developments include nanogels, which are able to encapsulate cisplatin. Such approaches showed reduced toxicity in comparison with the free drug, inhibited tumour growth, and promoted extended survival of C6 glioma-bearing rats when the nanogel was functionalized with a monoclonal antibody targeted to connexin 43, a protein highly expressed in the tumour periphery of C6 gliomas [175].
Since GB usually relapses after eventual transient response, and there is no validated second-line treatment, the co-encapsulation of synergistic drugs was proposed as a therapeutic alternative. It has been shown that cisplatin and fisetin encapsulated into liposomes have a synergistic effect, due to the combination of the anti-angiogenic effect of fisetin with the cytotoxic effect of cisplatin. The formulation showed an additive effect of cisplatin and fisetin against GB cells, demonstrating antitumoural effect [176]. Although combinatorial therapy based on TMZ plus cisplatin showed promising potential for GB therapy in clinical trials, the limited BBB crossing, poor targeting of GB tissues/cells, and systemic toxicity altogether hindered its efficacy in GB therapy. More recently, camouflaged GB cell membrane and pH-sensitive biomimetic nanoparticles demonstrated efficiently co-loading TMZ and cisplatin, providing transport across the BBB, and specifically targeting GB [177]. With this nanoformulation, a controlled release of drug cargos was obtained. A potent anti-GB effect in vivo was observed after treatment of mice bearing orthotopic U87 or the drug-resistant U251R GB tumours. The average survival was increased in mice receiving the combined drug administration compared to the survival time for mice receiving single-drug loaded nanoparticles, without noticeable side effects.
Apart from the development of novel nanoformulations, the understanding of drug transport across the BBB remains limited and represents a significant challenge for HGG treatment. There is an urgent need for predictive in vitro models with realistic and dynamic blood–brain barrier vasculature features. Thus, different models have been attempted to simulate the brain tumour vasculature such as microfluidic devices emulating BBB using self-assembled endothelial cells, astrocytes, and pericytes using modular layer-by-layer assembly [178]. The objective of such approaches is to investigate the transport of targeted nanotherapeutics across the BBB and into GB cells. In the reported study, cisplatin-loaded nanoparticles decorated with GB-targeting motifs to improve tumour trafficking were evaluated in this in vitro platform. The obtained results were compared with transport across mouse brain capillaries using intravital imaging, validating the ability of the platform to properly model the in vivo BBB transport. These models represent a significant advance, enabling in-depth investigation of brain tumour vasculature and accelerating the development of targeted nanotherapeutics.
The nanoformulation of cisplatin has demonstrated to remarkably improve its pharmacokinetics, which is especially unfavorable in case of glioblastomas, mainly due to challenges posed by biological barriers such as the BBB. However, the pharmacokinetic profile greatly depends on the type of nanostructure being used. In general, the integration of platinum complexes into nanoparticles allows for larger blood circulation time, more efficient arrival to the tumour zone while decreasing the amount of dosage needed, as well as protecting drugs from degradation, which increases stability. An example is a work published by Yu et al. that describes the pharmacokinetics of cisplatin-loaded polymeric nanoparticles [179]. The studies in vivo demonstrated that the selected nanoparticles had a long blood circulation time, the platinum concentration remained up to 46-fold higher than that of mice receiving equivalent doses of cisplatin, and the platinum concentration ratio of NPs to free cisplatin in tumours (lung tumour) reached as high as 9.4. Moreover, NPs improve the safety and tolerance in vivo, and improves the anticancer efficacy in comparison to cisplatin. Facts certainly change when talking about glioblastomas with the associated challenges for tumour delivery across the BBB: in this case, still, improvements can be obtained as shown by Ashrafzadeh et al., who reported a targeted pegylated liposomal cisplatin able to cause an increase in the cellular uptake and in brain tumour by 1.43 and 1.7-fold, respectively [168]. The administration of the NPs showed enhanced efficacy and reduced toxicity for the treatment of brain tumour. In fact, the blood–drug concentration was increased in comparison to cisplatin and then maintained in the effective range in a sustained mode. In general, the nanoformulations increase drug efficacy and low toxicity, and provide various pharmacokinetic benefits such as lowered drug accumulation with chronic drug dosing, and minimal fluctuation of drug concentration in blood.

5.2. Carboplatin

Carboplatin has a lower nephrotoxicity and ototoxicity incidence than cisplatin, although it is less active mainly due to its lower cellular uptake [180], thus repeated administration cycles with a large number of doses are required to inhibit tumour growth [181]. Amongst the platinum complexes (carboplatin, cisplatin and oxaliplatin), carboplatin exhibited the lowest toxicity while providing the best survival benefits. In addition, studies inducing DNA damage by low-energy secondary electrons produced by radiation suggest that carboplatin is better than cisplatin as a radiosensitizer [182]. These studies indicate that carboplatin is the best candidate among the approved platinum drugs for treatment of human brain tumours.
In an earlier study was described the covalent coupling of carboplatin to N-(2-Hydroxypropyl)methacrylamide (HPMA) and the tetrapeptide glycyl-phenylalanyl-leucyl-glycine (GFLG), whose proteolytic cleavage allows for its intracellular drug delivery [183,184]. The resulting nanoparticles revealed a higher therapeutic effect in preclinical models (compared to free carboplatin) and were accepted in a Phase I clinical trial [185]. On the other hand, the encapsulation of carboplatin with poly (ε-caprolactone) NPs was able to minimize hemolysis and increase the cytotoxicity effect against U87 human GB cells, which was more remarkable than the free drug [186]. Recently, it has been reported that the use of carboplatin-loaded poly (butyl cyanoacrylate) (PBCA) NPs conjugated with monoclonal antibodies against epidermal growth factor receptors (EGFR) for GB treatment. In this study, both therapeutic efficacy and side effects of the resulting NPs were evaluated in vivo in a rat model of GB (see Figure 4) [187]. The findings showed a 40% augmented cytotoxicity compared to the free form of carboplatin. Moreover, in vivo studies demonstrated an increased survival and appearance of less side effects in brain, kidney, or liver, compared to rats treated with free carboplatin. In another study, authors have described biodegradable PLGA NPs with controlled carboplatin release in rat brain by CED [188]. The treatment involved recurrent infusions due to the rapid clearance of carboplatin from the brain, and improved outcome was observed in GB patients, essentially due to an increase in cytotoxicity against tumours, lower neuronal toxicity, and prolonged half-life.
Although liposomes are the principal nanosystems studied for drug delivery, the most appropriate liposomal formulation for CED brain tumour injection remains to be determined. Different liposomal carboplatin formulations were prepared and tested in vitro in F98 glioma cells and in Fischer rats carrying orthotopic F98 tumours. The results indicated that the therapeutic efficacy in vitro and in vivo may vary drastically depending on surface charge [189]. Cationic liposomes bind more efficiently to tumour cells, increasing their therapeutic efficacy in vitro although the median survival time did not significantly improve. Interestingly, anionic and pegylated liposomes diffuse better towards the tumoural area, increasing its concentration in the tumour, reducing its clearance rate, and reducing neurotoxicity relative to free carboplatin. Thus, intratumour injection of liposomal carboplatin nanoformulations is considered a promising alternative to the administration of pure carboplatin in the chemotherapeutic treatment of GB.
A comparative study tested cisplatin, oxaliplatin, carboplatin, Lipoplatin (liposomal formulation of cisplatin), and LipoxalTM (liposomal formulation of oxaliplatin) administered by intracarotid infusion to F98 gliomas orthotopically growing in Fischer rats. The nanoformulation of platinum compounds demonstrated reduced toxicity, improved cancer cell uptake, and increased survival rates either when combined or not with radiotherapy, in comparison with liposome-free drugs. Moreover, among the platinum compounds tested, liposomal carboplatin showed the largest survival increase when combined with radiation in vivo [97].
Interestingly, in the search of new routes of administration, Alex et al. encapsulated carboplatin into polycaprolactone NPs to target GB via the nasal route [190]. The NPs showed improved in vitro anti-tumour activity in comparison with the free drug against the human GB cell line LN229. Ex vivo permeation studies through sheep nasal mucosa showed a similar release pattern as for in vitro release studies. Moreover, in situ nasal perfusion administration in Wistar rats demonstrated that NPs show better nasal absorption than carboplatin solution while no severe damage to the integrity of nasal mucosa was detected. These results suggest that these nanoformulations are suitable to improve the nasal absorption of carboplatin and consequently to improve brain delivery.

5.3. Oxaliplatin

Oxaliplatin was demonstrated to induce multi-faceted anti-tumour effects, more effectively than cisplatin and carboplatin at drug concentrations/doses below those required to induce apoptosis, which fostered specific research in the GB area. However, in vivo preclinical studies were unsatisfactory since i.v. administration did not increase the median survival time of F98 glioma-bearing rats [104]. This may be most likely due to the BBB preventing drug accumulation in brain tumours, even when compromised such as in HGG [191]. To overcome these limitations, the drug was bound to PEG-glutamic acid (Glu) micelles functionalized with cyclic Arg-Gly-Asp (cRGD) in order to cross the BBB, target GB cells, and penetrate into GB via cRGD-mediated transvascular transport (see Figure 5) [192]. Their antitumour effect against GB was compared to the effect of control NPs functionalized with the nontargeted peptide. These nanoformulations exhibited significant growth inhibition effects against GB compared to the non-targeted micelles, as well as tumour accumulation, indicating the active transport of cRGD-mediated drug delivery across vascular and tumour barriers.
Later, the liposomal formulation containing oxaliplatin (Lipoxal™) was administered via CED, which successfully improved tumour accumulation and subsequent survival time in F98 glioma-bearing rats when combined with radiotherapy. Moreover, the maximum tolerated dose (MTD) of Lipoxal™ was 3-fold superior to that of free oxaliplatin [193]. More recently, You et al. reported multiwalled carbon nanotubes (MWCNTs) containing oxaliplatin functionalized with cell penetrating peptides (TAT) and the cancer targeting molecule biotin. This nanosystem enhanced oxaliplatin cytotoxicity towards glioma cells as a result of reactive oxygen species (ROS) overproduction while in vivo studies demonstrated its BBB penetration and outstanding antitumour efficacy against orthotopic glioma [194].
Apart from the gold-standard platinum complexes, there are some additional developments of novel platinum-based nanoformulations with interesting results concerning GB treatment. This is the case of a reported platinum-based polyethylenimine (PEI) polymer–drug conjugate with potential anti-GB stem cells. The cationic polymer presented a potent and specific toxicity in vitro. The results obtained from cytotoxicity studies with NCH421K and NCH644 GN cells indicated a necrotic cell death mechanism with an absence of apoptotic markers. Several markers also indicated that this cell death mechanism could induce an anti-cancer immune response [195].

5.4. Pt(IV) Prodrugs

Pt(IV) prodrugs have been used mainly to counteract cisplatin resistance and nephrotoxicity effects; while in its Pt(IV) oxidation state, side effects are considerably minimized, though inside the cells it is further reduced to the antitumoural active Pt(II) form [196]. Thanasupawat et al. reported self-assembled coiled nanotubes linked with a Pt(IV) complex that exhibit better in vitro and in vivo toxicity against human U87 GB cells (and cells obtained from human GB samples) than free Pt(IV), mostly due to the activation of multiple death pathways. Moreover, the nanosystems were active in subcutaneous/orthotopic U87 xenografts using intratumoural administration [93].
A more complete study from Dhar and coworkers in 2015 reported a lipophilic polymeric NP carrying a Pt(IV)-prodrug (see Figure 6) with capabilities of mitochondria targeting [128]. The platinum prodrug of cisplatin (Platin-M) was encapsulated in targeted polymeric nanoparticles carrying the drug across the BBB and specifically towards mitochondria. NP-mediated controlled release of Platin-M and its subsequent reduction to cisplatin provoked the cross-linking with the mitochondrial DNA, thus forcing overactive cancer cells to undergo apoptosis. In vitro effects of the NPs in canine glioma and GB cell lines demonstrated to be much more effective than free cisplatin or carboplatin. In vivo biodistribution studies in healthy adult beagles after single i.v. injection showed high levels of Pt accumulation into the brain, with negligible amounts found in other organs. Moreover, no signs of neurotoxicity were observed, demonstrating the translational potential of these nanoformulation for applications in brain tumour treatment.
A recent study showed that NPs encapsulating an oxaliplatin prodrug and a cationic DNA intercalator, administered through CED, were able to inhibit the growth of TMZ-resistant cells from patient-derived xenografts, and hinder the progression of TMZ-resistant human GB tumours in mice, without causing any noticeable toxicity [197]. Such polymeric nanoformulations contain disulfide bonds which are cleaved in the reductive tumour environment, which affords a selective release of anticancer drugs. The reported research findings suggest that the administration of drugs with different mechanisms of action, combined with CED, may represent a translational strategy for the treatment of TMZ-resistant gliomas. Another example includes polyethylene glycol (PEG)-stabilized solid lipid nanoparticles (SLNs) containing different Pt(IV) prodrugs derived from kiteplatin. An in vitro BBB model of immortalized human cerebral microvascular endothelial cells (hCMEC/D3) was used for evaluating the ability of the SLNs to cross the BBB, while U87 human GB was used for cell internalization and antitumour efficacy studies. These nanoformulations demonstrated potent ability to permeate the in vitro BBB model with improved cellular uptake and higher reduction in cell viability in comparison with their free counterparts [198].
A novel type of nanoparticles containing Pt(IV) prodrugs as building block of nanostructured coordination polymers (NCPs) was recently described, with enormous potential for treating different cancers [157]. The versatility of coordination chemistry allows obtaining nanoparticles by reaction of a novel platinum (IV) prodrug and metal ions (i.e., iron or zinc) producing theranostic nanosystems [199]. Ruiz-Molina et al. designed a nanoparticle based on a coordination polymer with building blocks containing Pt(IV) complexes obtained from cisplatin and iron ions as metallic nodes. This nanoformulation presents dual pH and redox sensitivity in vitro, showing controlled release and comparable cytotoxicity to cisplatin against HeLa and GL261 GB cells. Interestingly, in vivo intranasal administration in orthotopic preclinical GL261 GB-bearing mice demonstrated increased accumulation of platinum in tumours, leading in certain cases to cure and prolonged survival of the tested cohort (Figure 7) [96]. Additionally, the presence of magnetically active iron(III) ions allows these NPs to be proposed as potential nanoprobes to be tracked in vivo by MRI, thus opening new opportunities for the design of intranasal glioblastoma therapies while minimizing side effects.
Table 1. Preclinical studies on the effect of nanoformulated anticancer Pt-complexes for GB treatment.
Table 1. Preclinical studies on the effect of nanoformulated anticancer Pt-complexes for GB treatment.
Pt AgentNanocarrier
& Size
FormCells/Animal Models
(Administration Mode)
Effect/Mechanism of ActionRef.
CisplatinIron oxide NPs in NGO
(lateral width of 80–100 nm and thickness of 6.3 nm)
γ-Fe2O3 NPs coated by nanographene oxide (NGO) containing cisplatinU87 cell lineSuperparamagnetic-like behavior but showed little cytotoxicity at 10 μM.[159]
CisplatinAu-NPs
(7 nm)
Cisplatin conjugated to gold NPsU251/U87 cell lines, U251 xenografts (intravenous injection)Inhibited GB cells growth and showed marked synergy with MR-guided Focused Ultrasound (MRgFUS) therapy. In vivo assays demonstrated increased BBB permeability.[160]
CisplatinAu-NPs
(145 nm)
Cisplatinum attached to hyaluronic acid-functionalized gold NPs and coated with PEGU87 cell line
MCF-7
tumour-bearing mice
(intravenous injection)
NPs inhibited GB cells growth and showed marked synergy with the thermal effect of applying NIR laser radiation.[161]
CisplatinLiposomes
(50–60 nm)
LipoplatinTM and novel cisplatin bound to coordination complexes with lipid cholesteryl hemisuccinate (CHEMS)F98 glioma cell line
F98 glioma-bearing rats
(convection-enhanced delivery)
Potent in vitro cytotoxicity. High intracerebral retention after CED. CHEMS liposomes were better tolerated than LipoplatinTM and showed increased survival data in vivo.[165]
CisplatinLiposomes
(130–160 nm)
Cisplatin analogue and antibodies against VEGF or VEGFR2 conjugated on liposome surface C6, U87 cell linesProlonged blood circulation time in C6 glioma-bearing rats. In vitro data confirmed that conjugation with specific antibodies increases the intracellular concentration of the liposomes and improves cytotoxicity.[167]
CisplatinLiposomes
(160 nm)
Liposomes decorated with OX26 monoclonal antibodies C6 cell line
C6 glioma-bearing rats
(stereotactic injection)
Increased cellular uptake and increased mean survival time in vivo with notable reduction in toxicity effects.[168]
CisplatinPolymeric NPs
(105 nm)
Protamine-functionalized PLGA NPs U87 cell lineIncreased cellular uptake and cytotoxicity in vitro and ability to cross in vitro BBB model, improving therapeutic index.[169]
CisplatinPolymeric NPs
(50–145 nm)
PLGA-PEG copolymer NPs functionalized with a triphenylphosphonium cation C6 cell line
C6 glioma-bearing rats
(intravenous injection)
Nanoparticles below 100 nm diffused within mice brain if densely coated with PEG.[170]
CisplatinPolymeric NPs
(114 nm)
PLGA NPsU87 cell lineSignificantly higher cytotoxicity than cisplatin, enhanced cellular uptake.[171]
CisplatinPolymeric NPs
(490 nm)
PBCA NPsC6 cell line
C6 glioma-bearing rats
(intraperitoneal injection)
Similar cytotoxicity in comparison to cisplatin. Slightly longer mean survival time and reduced side effects.[172]
CisplatinPolymeric NPs
(70 nm)
PEGylated-Poly(aspartic acid) NPs9L gliosarcoma and F98 glioma lines, orthotopic F98 brain tumour model (convection-enhanced delivery)Reduction in toxicity associated with cisplatin and increased average survival in vivo.[173]
CisplatinPolymeric NPs
(9 nm)
Gd-grafted PEO micellesU87/U251 cell lines,
U87 orthotopic xenografts
(stereotactic injection)
Hyperintense signal on T2-weighted MRI, 50-fold increased cellular accumulation, and 32-fold Pt-DNA adduct compared to free cisplatin.[174]
CisplatinNanogel
(110 nm)
PEG-b-PMAA nanogels functionalized with monoclonal antibody to Cx43C6 cells
C6 glioma-bearing rats
(intravenous injection)
Targeted delivery to C6 glioma. Prolonged the mean
survival time of glioma-bearing rats.
[175]
Cisplatin + Glutathione
Peroxidase
Iron NPs
(80–125 nm)
Cisplatin and small
interfering RNA (siRNA) targeting glutathione peroxidase attached to IONPs and functionalized with folic acid
U87 cell line, primary glioblastoma cell line (P3#GB), normal human
astrocytes (NHAs)
U87orthotopic xenografts (stereotactic injection)
Potent ROS and ferroptosis, synergistically improved therapeutic efficacy, low systemic toxicity achieved both in vitro and in vivo. [86]
Cisplatin + FisetinLiposomes
(173 nm)
Both drugs incorporated to a liposomal formulation composed by DOPC/cholesterol/DODA-GLY-PEG2000U87 cell lineAdditive effect of cisplatin and fisetin. Effective antitumoural action against GB cells.[176]
Cisplatin + TMZBiomimetic NPs
(171 nm)
pH degradable acetal grated dextran
inner core coated with GB cancer cell membrane
U87 or drug-resistant U251R GB tumours
(intravenous injection)
Improvement in BBB penetration and targeting of GB tissue/cells. Potent anti-GB activity in mice bearing orthotopic U87 or TMZ resistant U251 (U251R) tumours.[177]
CarboplatinPolymeric NPsPolymer poly (ε-caprolactone) NPsU87 MGIncrease uptake and cytotoxicity in U87 human GB cell line. No haemolytic activity in rat erythrocytes.[186]
CarboplatinPolymeric NPs
(20–100 nm)
Polymer poly (ε-caprolactone) NPsLN229 human GB cell lineImproved in vitro anti-tumour activity in comparison to free Carboplatin. Better nasal absorption compared to pure carboplatin solution and no severe damage of nasal mucosa.[190]
CarboplatinPolymeric NPs
(365 nm)
Carboplatin-loaded poly (butyl cyanoacrylate) (PBCA) NPs conjugated with monoclonal antibodyC6 glioma cell lines
C6 glioma-bearing rats
(intraperitoneal injection)
Augmented cytotoxicity compared to free carboplatin. In vivo studies demonstrated an increased survival time and less side effects in the brain, kidney, or liver, compared to free carboplatin.[187]
CarboplatinPolymeric NPs
(200 nm)
PLGA NPsPrimary rat hippocampal cell cultures.
Adult male Wistar rats and white pigs
(convection-enhanced delivery)
Improved cytotoxicity in vitro.
Negligible neurotoxic effects and prolonged tissue half-life in vivo after infusion by CED in small and big animal models.
[188]
OxaliplatinPolymeric NPs
(20–30 nm)
PEG-P(Glu) NPs functionalized with cyclic Arg-Gly-Asp (cRGD)U87MG cell lineTarget GB cells and ability to cross BBB. Tumour growth inhibition in orthotopic model.[192]
OxaliplatinLiposomes
(57–86 nm)
Liposomal formulation (LipoxalTM)F98 cell line
(convection-enhanced delivery)
Administered by CED, increased tumoural accumulation, and median survival time in F98 glioma-bearing rats.[193]
OxaliplatinCNTs
(20 nm–several μm)
Multi-walled carbon nanotube functionalized with TAT/BiotinC6 and CHEM-5 cell lines, C6 orthotopic glioma
(intravenous injection)
Improved cytotoxicity of oxaliplatin toward glioma cells. Enhanced BBB penetration, improved brain-targeting selectivity, and excellent anti-tumour activity against orthotopic glioma. [194]
Pt(IV) prodrugPolymeric NPs
(50–145 nm)
Cisplatin Pt(IV) prodrug in PLGA–PEG NPsCanine J3TBG glioma and SDT3G glioblastoma cell lines
(intravenous injection)
Overall 17 times more effective than cisplatin in vitro.
Capable of crossing the canine BBB and accumulating in the brain.
[128]
Pt(IV) prodrugCNTs
(20 kDa ̴ 4 nm)
Coil nanotubesU87, U251, and patient GB cells, human astrocytes, U87 xenografts
(subcutaneous injection)
Enhanced in vitro and in vivo tumour cell killing in comparison with free prodrug, activating multiple cell death pathways in GB cells without affecting astrocytes in vitro or causing damage to healthy mouse brain.[93]
Pt(IV) prodrugPolymeric NPs
(105 nm)
Oxaliplatin-derived Pt(IV) complexes attached to a reduction-responsive polymer,
(poly(CHTA-co-HD)-PEG)
U87, TMZ-resistant
LN229 cell lines and cells from a patient-derived xenograft.
In vivo LN229-TR xenografts
(convection-enhanced delivery)
Improved cell uptake and cytotoxicity in comparison to TMZ.
Lack of toxicity, favorable distribution of drugs into the region of interest, substantial tumour growth inhibition, and increased by 3 the survival time in mice bearing tumours after CED administration in vivo.
[197]
Pt(IV) prodrugsSolid lipid NPs (SLNs)
(30–80 nm)
Pt(IV) prodrugs derived from kiteplatin in (SLNs)U87 human GB cell line, hCMEC/D3 endothelial cellsEnhanced permeability, improved cell uptake compared to free drug.[198]
Pt(IV) prodrugsNanostructured Coordination Polymers (NCPs)
(70 nm)
Pt(IV) prodrugs obtained from cisplatin as building block in NCPsHeLa and GL261 cell lines.
Orthotopic preclinical GL261 GB tumour-bearing mice
(intranasal administration)
In vitro dual pH and redox-mediated control release.
Comparable cytotoxicity to cisplatin.
In vivo intranasal administration demonstrated increased tumour accumulation of platinum and effective anticancer action.
[199]
Pt0 NPsMetal NPs
(2–20 nm)
FePt NPsU251, U87, and H4 cell linesMarked cytotoxicity observed in lipophilic coated FePt NPs and low cytotoxicity in the case of hydrophilic FePt NPs.[200]
Pt(0)Metal NPs
(42 nm)
Ag-Pt NPsU87 cell lineInhibition of Gram-positive and Gram-negative multi-drug resistant bacteria. Selective and dose-dependent anticancer activity over U87 GB cells without being harmful to healthy human fibroblasts.[201]
Some references concerning use of metal-based Pt nanoparticles for GB treatment can also be found in literature. Thus, Kutwin et al. reported a comparative study between Pt-NPs and cisplatin against U87 GB cells or GB tumours growing on chorioallantoic membranes, observing that NPs showed antiproliferative activity although it was significantly lower than cisplatin [200]. Additionally, Lopez Ruiz et al. reported AgPt NPs with a selective and dose-dependent anticancer activity on human U87 GB and A375 melanoma cell lines (10–250 μg/mL concentration range) without compromising the activity of healthy human fibroblasts [201].
Concerning the best route of administration for nanoformulations containing platinum-based drugs, Paquette et al. published a series of papers combining Pt-based complexes and radiotherapy while assessing different administration routes, being the best results achieved with intra-arterial administration [104]. The same authors had already reported in 2012 cisplatin (lipoplatin) or oxaliplatin (lipoxal) liposomes administered by intracarotid infusion in orthotopic F98 glioma-bearing rats, obtaining notable results related to reduced toxicity, improved cell uptake, and increased survival of animals when combined with radiotherapy [97]. Afterwards, they also reported the combination of Pt drugs (administered via CED) and radiation treatment in the same murine model, increasing tumour drug retention, reducing side effects, and achieving a larger median survival time [202]. The combination of cisplatin-bearing AuNPs with radiotherapy was also used in in vitro resistant conditions to enhance DNA double-strand rupture and therefore apoptosis rate [203].

6. Why Repurpose Platinum Drugs and Use Nanotechnology?

Treatment of GB has remained almost unchanged for more than 20 years despite of continuous research efforts towards improvement. In the search of new treatment, platinum-based drugs have resurfaced as suitable alternatives for GB therapy. Although the developed anticancer platinum drugs are being used successfully to treat a variety of cancer types, most clinical assays with GB patients fail, in general due to dose-limiting toxicities when delivered systemically or in the tumour region, the low amount of drug doses that crosses the BBB, and/or systemic toxicities occurring before effective drug concentrations can reach tumours. However, the relative success of platinum drugs for treatment of non-CNS cancers suggests that there is hope regarding their therapeutic potential, provided they are properly delivered to the tumour region. Moreover, additional effects of platinum drugs such as immunomodulatory effects, and the use of specific delivery strategies that can maximize these multimodal effects and minimize toxicities, may help to foster their re-purposing. Early clinical trials with these agents offered great promise by using cisplatin [204], carboplatin [205], or oxaliplatin [206,207]. However, in most of the clinical trials, the combination of platinum drugs with radiation therapy and/or other chemotherapeutic agents showed no significant survival advantage. In addition, systemic and dose-limiting toxicity remains a key challenge related to the administration of platinum chemotherapeutics, including nanoformulations such as LipoplatinTM which is currently in a Phase III clinical trial for treatment of non-small cell lung cancer.
The DNA damaging effect of platinum-based drugs affects pathways related to cell invasion, angiogenesis, chemo- and radio-sensitization, and immunomodulation. A lower sustained platinum drug dose over longer times permits extended cell viability and reorganization of complex cellular pathways and the tumour microenvironment. Platinum compounds can considerably reduce the ability of GB cells to invade the surrounding tissue by downregulating matrix metalloproteinase (MMP) expression [208]; this is one of the key advantages of repurposing platinum drugs for GB therapy. Moreover, platinum compounds may have anti-angiogenic effects [41] and can enhance the efficacy of the current adjuvant therapies for GB (i.e., TMZ and radiotherapy) [42,43].
Additionally, it has been demonstrated that platinum drug-related immunomodulation can have an incidence in reversing GB-mediated immune evasion [209]. Moreover, platinum drugs are capable of modulating immunosuppressive features associated with numerous cancers including GB [210,211] and could even alter the profile of circulating immune cells or tumour-infiltrating immune cells [212]. Specifically, it has been demonstrated that oxaliplatin can induce immunogenic cell death, enhancing antitumour adaptive immune responses against antigens expressed by the dead cells [213]. However, cisplatin does not seem to induce immunogenic cell death [211]. Thus, future studies would be needed in order to investigate which platinum drugs are capable of modulating the glioma microenvironment, and to which extent, to improve the outcome of cancer treatments.
Apart from the intrinsic effect of the platinum complexes, the use of nanoformulated systems may also foster improvement of the therapeutic outcomes. As it has been shown through different examples along this review, the integration of platinum drugs and prodrugs in a nanocarrier permits to protect the active compound and increase its delivery to the desired sites. Moreover, it can improve platinum-based therapies by (i) increasing the solubility of the complexes and their blood half-life, (ii) reducing side effects through targeted delivery and broader tissue distribution, (iii) providing a sustained drug release, and (iv) allowing simultaneous incorporation and delivery of other anticancer drugs for combination therapy [214].

7. Summary and Future Perspectives

The scarce advancements in GB therapy and lack of efficient therapeutic alternatives makes this disease a relevant field for pushing forward cutting-edge research. In this regard, nanoencapsulation can offer a new dimension for the assessment of both novel drugs and classic compounds with poor BBB penetration and/or previous failure in clinical trials. Among the already known drugs with demonstrated anticancer activity, but not included as usual therapeutic option at GB relapsing, platinum complexes have properties that make them especially suitable for the development of novel nanotechnology-based therapeutics. Cisplatin is usually chosen as frontline treatment [35,173], while carboplatin is rather used at tumour recurrence [215]. Recent studies using platinum drugs or prodrugs, alone or combining cisplatin with other adjuvants, have proven promising improvement in survival rates. Overall, these factors might place the platinum complexes back in the spotlight to test the new pharmacokinetic advantage paradigms provided by the nanoparticle technology. However, even though successful preclinical studies have already been reported, no clinical phases have been reached, preventing products to reach the market.
The previously unrecognized or underestimated potential of platinum-based treatments can now offer another opportunity for platinum complexes. By using the principles of nanomedicine and chemistry formulation, new approaches for GB treatment can be developed, and hopefully reach clinical phases in the future. NPs can overcome some of the drug delivery limitations posed by the BBB, providing a controlled and sustained drug release in the site of action, targeting tumour cells, and reducing toxicity. Worth to mention, immunomodulation might be included in the newly recognized therapeutic effects, which may have broad application in future combination therapies for GB. In recent years, immunotherapy strategies have revolutionized the treatment of many cancer types, also increasing the hope for GB therapy. However, mostly due to the multifactorial immunosuppression occurring in the GB microenvironment, the poor knowledge of the neuroimmune system, and the presence of the BBB, the efficacy of immunotherapy in GB is still limited. New and recent strategies for GB treatment have employed immunotherapy combinations and have provided encouraging results in both preclinical and clinical studies. The lessons learned from clinical trials highlight the importance of tackling different arms of immunity.
Another interesting aspect lies in the investigation of novel non-invasive administration routes such as intranasal delivery, as alternatives to classical routes (i.e., CED or intravenous administration). Non-invasive brain targeting through the intranasal route arises as an attractive modality to circumvent the BBB along the trigeminal/olfactory neuronal pathways, thereby enhancing the therapeutic ratio with minimal systemic exposure [216,217]. The main goal of such novel administration routes is to provide effective drug concentration in brain tumours using reduced doses, overcoming the BBB while avoiding most of the systemic side effects. In other words, it is not a matter of how much Pt reaches the tumour site, but ensuring that enough Pt reaches the tumour milieu in order to trigger immunogenic cell damage and elicit host immune system. Last, but not least, it is also relevant to avoid systemic effects that are a limiting factor for single and combined Pt treatments for GB.
Additionally, besides the therapeutic use of Pt complexes as single drugs, the combination of Pt-based nanoparticles with other approaches such as chemotherapy, radiotherapy, immunotherapy, gene therapy, or functional inhibitors seems to be attractive and raise interesting possibilities. In this sense, it is important to highlight the preclinical evidence regarding combination immunotherapy in terms of both immune and survival benefits for GB management. Recent studies assessing the combination of different classes of immunotherapeutic agents (e.g., immune checkpoint blockade and vaccines) offer interesting perspectives that may be an inspiring starting point to develop future strategies facilitating the clinical translation to address the unmet medical needs of GB treatment.

Author Contributions

F.N., J.L., A.P.C. and D.R.-M. contributed to the conceptualization, methodology, validation, literature search, writing original, draft preparation, writing review and editing. C.A. and P.A.-T. contributed to the draft preparation, study design, literature search, data collection, data analysis, and data interpretation. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by grant PID2021-127983OB-C21, PID2021-127983OB-C22, and PID2020-113058GB-I00 funded by MCIN/AEI/10.13039/501100011033 and by ERDF A way of making Europe and by PI21/00181 from the Instituto Carlos III. The ICN2 is funded by the CERCA programme/Generalitat de Catalunya. The ICN2 is supported by the Severo Ochoa Centres of Excellence programme, Grant CEX2021-001214-S, funded by MCIN/AEI/10.13039.501100011033. This work was also supported by the Centro de Investigación Biomédica en Red-Bioingeniería, Biomateriales y Nanomedicina (CIBER-BBN [http://www.ciber-bbn.es/en, accessed on 10 April 2023], CB06/01/0010), an initiative of the Instituto de Salud Carlos III (Spain).

Data Availability Statement

No new data were created.

Acknowledgments

This work was also supported by “Metalfármacos multifuncionales para el diagnóstico y la terapia” network, through grant RED2018-102471-T, funded by MCIN/AEI/10.13039/501100011033. COST is kindly acknowledged for stimulating scientific discussion via the Cost Action CA 17140 “Cancer nanomedicine from the bench to the bedside”.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hanif, F.; Muzaffar, K.; Perveen, K.; Malhi, S.M.; Simjee, S.U. Glioblastoma Multiforme: A Review of its Epidemiology and Pathogenesis through Clinical Presentation and Treatment. Asian Pac. J. Cancer Prev. 2017, 18, 3–9. [Google Scholar]
  2. Ostrom, Q.T.; Cote, D.J.; Ascha, M.; Kruchko, C.; Barnholtz-Sloan, J.S. Adult Glioma Incidence and Survival by Race or Ethnicity in the United States from 2000 to 2014. JAMA Oncol. 2018, 4, 1254–1262. [Google Scholar] [CrossRef]
  3. Chen, X.; Zhang, M.; Gan, H.; Wang, H.; Lee, J.-H.; Fang, D.; Kitange, G.J.; He, L.; Hu, Z.; Parney, I.F.; et al. A novel enhancer regulates MGMT expression and promotes temozolomide resistance in glioblastoma. Nat. Commun. 2018, 9, 2949. [Google Scholar] [CrossRef]
  4. Broekman, M.L.; Maas, S.L.N.; Abels, E.R.; Mempel, T.R.; Krichevsky, A.M.; Breakefield, X.O. Multidimensional communication in the microenvirons of glioblastoma. Nat. Rev. Neurol. 2018, 14, 482–495. [Google Scholar] [CrossRef]
  5. Brown, N.F.; Carter, T.J.; Ottaviani, D.; Mulholland, P. Harnessing the immune system in glioblastoma. Br. J. Cancer. 2018, 119, 1171–1181. [Google Scholar] [CrossRef]
  6. Chen, D.S.; Mellman, I. Oncology meets immunology: The cancer-immunity cycle. Immunity 2013, 9, 1–10. [Google Scholar] [CrossRef]
  7. Condoluci, A.; Mazzara, C.; Zoccoli, A.; Pezzuto, A.; Tonini, G. Impact of smoking on lung cancer treatment effectiveness: A review. Future Oncol. 2016, 12, 2149–2161. [Google Scholar] [CrossRef]
  8. Stupp, R.; Warren, P.; Mason, P.; van den Bent, M.J.; Weller, M.; Fisher, B.; Taphoorn, M.J.B.; Belanger, K.; Brandes, A.A.; Marosi, C.; et al. Radiotherapy plus Concomitant and Adjuvant Temozolomide for Glioblastoma. N. Engl. J. Med. 2005, 352, 987–996. [Google Scholar] [CrossRef]
  9. Goldbrunner, R.; Ruge, M.; Kocher, M.; Lucas, C.W.; Galldiks, N.; Grau, S. The Treatment of Gliomas in Adulthood. Dtsch. Arztebl. Int. 2018, 115, 356–364. [Google Scholar] [CrossRef]
  10. Mowforth, O.D.; Brannigan, J.; El Khoury, M.; Sarathi, C.I.P.; Bestwick, H.; Bhatti, F.; Mair, R. Personalised therapeutic approaches to glioblastoma: A systematic review. Front. Med. 2023, 10, 1166104. [Google Scholar] [CrossRef]
  11. Messali, A.; Villacorta, R.; Hay, J.W. A Review of the Economic Burden of Glioblastoma and the Cost Effectiveness of Pharmacologic Treatments. PharmacoEconomics 2014, 32, 1201–1212. [Google Scholar] [CrossRef]
  12. Reid, J.M.; Stevens, D.C.; Rubin, J.; Ames, M.M. Pharmacokinetics of 3-methyl-(triazen-1-yl) imidazole-4-carboximide following administration of temozolomide to patients with advanced cancer. Clin. Cancer Res. 1997, 3, 2393–2398. [Google Scholar]
  13. Lee, S.Y. Temozolomide resistance in glioblastoma multiforme. Genes Dis. 2016, 3, 198–210. [Google Scholar] [CrossRef]
  14. Alonso, M.M.; Gomez-Manzano, C.; Bekele, B.N.; Yung, W.A.; Fueyo, J. Adenovirus-based strategies overcome temozolomide resistance by silencing the O6-methylguanine-DNA methyltransferase promoter. Cancer Res. 2007, 67, 11499–11504. [Google Scholar] [CrossRef]
  15. Montaldi, A.P.; Godoy, P.R.; Sakamoto-Hojo, E.T. APE1/REF-1 down-regulation enhances the cytotoxic effects of temozolomide in a resistant glioblastoma cell line. Mutat. Res. Genet. Toxicol. Environ. Mutagen. 2015, 793, 19–29. [Google Scholar] [CrossRef]
  16. Ramirez, Y.P.; Weatherbee, J.L.; Wheelhouse, R.T.; Ross, A.H. Glioblastoma multiforme therapy and mechanisms of resistance. Pharmaceuticals 2013, 6, 1475–1506. [Google Scholar] [CrossRef]
  17. Alphandéry, E. Glioblastoma Treatments: An Account of Recent Industrial Developments. Front. Pharmacol. 2018, 9, 879. [Google Scholar] [CrossRef]
  18. Rominiyi, O.; Vanderlinden, A.; Clenton, S.J.; Bridgewater, C.; Al-Tamimi, Y.; Collis, S.J. Tumour treating fields therapy for glioblastoma: Current advances and future directions. Br. J. Cancer. 2021, 124, 697–709. [Google Scholar] [CrossRef]
  19. Nam, J.Y.; de Groot, J.F. Treatment of Glioblastoma. J. Oncol. 2017, 13, 629–638. [Google Scholar] [CrossRef]
  20. Diaz, R.J.; Ali, S.; Qadir, M.G.; De La Fuente, M.I.; Ivan, M.E.; Komotar, R.J. The role of bevacizumab in the treatment of glioblastoma. J. Neurooncol. 2017, 133, 455–467. [Google Scholar] [CrossRef]
  21. Taylor, T.E.; Furnari, F.B.; Cavenee, W.K. Targeting EGFR for Treatment of Glioblastoma: Molecular Basis to Overcome Resistance. Curr. Cancer Drug Targets 2012, 12, 197–209. [Google Scholar] [CrossRef]
  22. Kreisl, T.N.; Kotliarova, S.; Butman, J.A.; Albert, P.S.; Kim, L.; Musib, L.; Thornton, D.; Fine, H.A. A phase I/II trial of enzastaurin in patients with recurrent high-grade gliomas. Neuro Oncol. 2010, 12, 181–189. [Google Scholar] [CrossRef]
  23. Heffron, T.P.; Ndubaku, C.O.; Salphati, L.; Alicke, B.; Cheong, J.; Drobnick, J.; Edgar, K.; Gould, S.E.; Lee, L.B.; Lesnick, J.D.; et al. Discovery of Clinical Development Candidate GDC-0084, a Brain Penetrant Inhibitor of PI3K and mTOR. ACS Med. Chem. Lett. 2016, 7, 351–356. [Google Scholar] [CrossRef]
  24. Medikonda, R.; Dunn, G.; Rahman, M.; Fecci, P.; Lim, M. A review of glioblastoma immunotherapy. J. Neuro-Oncol. 2021, 151, 41–53. [Google Scholar] [CrossRef]
  25. Liu, Y.; Wang, W.; Zhang, D.; Sun, Y.; Li, F.; Zheng, M.; Lovejoy, D.B.; Zou, Y.; Shi, B. Brain co-delivery of first-line chemotherapy drug and epigenetic bromodomain inhibitor formultidimensional enhanced synergistic glioblastoma therapy. Exploration 2022, 2, 20210274. [Google Scholar] [CrossRef]
  26. Weenink, B.; French, P.J.; Sillevis Smitt, P.A.E.; Debets, R.; Geurts, M. Immunotherapy in Glioblastoma: Current Shortcomings and Future Perspectives. Cancers 2020, 12, 751. [Google Scholar] [CrossRef]
  27. Huang, B.; Yu, Z.; Liang, R. Effect of long-term adjuvant temozolomide chemotherapy on primary glioblastoma patient survival. BMC Neurol. 2021, 21, 424. [Google Scholar] [CrossRef]
  28. Aliyu, A.; Shaari, M.R.; Sayuti, N.S.A.; Reduan, M.F.H.; Sithambaram, S.; Noordin, M.M.; Shaari, K.; Hamzah, H. N-Ethyl-n-Nitrosourea Induced Leukaemia in a Mouse Model through Upregulation of Vascular Endothelial Growth Factor and Evading Apoptosis. Cancers 2020, 12, 678. [Google Scholar] [CrossRef]
  29. Jeon, J.; Lee, S.; Kim, H.; Kang, H.; Youn, H.; Jo, S.; Youn, B.; Kim, H.Y. Revisiting Platinum-Based Anticancer Drugs to Overcome Gliomas. Int. J. Mol. Sci. 2021, 22, 5111. [Google Scholar] [CrossRef]
  30. Ghosh, D.; Nandi, S.; Bhattacharjee, S. Combination therapy to checkmate Glioblastoma: Clinical challenges and advances. Clin. Trans. Med. 2018, 7, 33. [Google Scholar] [CrossRef]
  31. Kelland, L. The resurgence of platinum-based cancer chemotherapy. Nat. Rev. Cancer 2007, 7, 573–584. [Google Scholar] [CrossRef] [PubMed]
  32. de Biasi, A.R.; Villena-Vargas, J.; Adusumilli, P.S. Cisplatin-Induced Antitumor Immunomodulation: A Review of Preclinical and Clinical Evidence. Clin. Cancer Res. 2014, 20, 5384–5391. [Google Scholar] [CrossRef] [PubMed]
  33. Ndagi, U.; Mhlongo, N.; Soliman, M.E. Metal complexes in cancer therapy—An update from drug design perspective. Drug. Des. Devel. Ther. 2017, 11, 599–616. [Google Scholar] [CrossRef]
  34. Wang, J.L.; Barth, R.F.; Cavaliere, R.; Puduvalli, V.K.; Giglio, P.; Lonser, R.R.; Elder, J.B. Phase I trial of intracerebral convectionenhanced delivery of carboplatin for treatment of recurrent high-grade gliomas. PLoS ONE 2020, 15, e0244383. [Google Scholar] [CrossRef] [PubMed]
  35. Silvani, A.; Eoli, M.; Salmaggi, A.; Lamperti, E.; Maccagnano, E.; Broggi, G.; Boiardi, A. Phase II trial of cisplatin plus temozolomide, in recurrent and progressive malignant glioma patients. J. Neuro-Oncol. 2004, 66, 203–208. [Google Scholar] [CrossRef] [PubMed]
  36. Jacobs, S.; McCully, C.L.; Murphy, R.F.; Bacher, J.; Balis, F.M.; Fox, E. Extracellular fluid concentrations of cisplatin, carboplatin, and oxaliplatin in brain, muscle, and blood measured using microdialysis in nonhuman primates. Cancer Chemother. Pharmacol. 2010, 65, 817–824. [Google Scholar] [CrossRef]
  37. Johnstone, T.C.; Suntharalingam, K.; Lippard, S.J. The Next Generation of Platinum Drugs: Targeted Pt(II) Agents, Nanoparticle Delivery, and Pt(IV) Prodrugs. Chem. Rev. 2016, 116, 3436–3486. [Google Scholar] [CrossRef] [PubMed]
  38. Huncharek, M.; Muscat, J. Treatment of recurrent high grade astrocytoma; results of a systematic review of 1415 patients. Anticancer Res. 1998, 18, 1303–1311. [Google Scholar]
  39. Yung, W.K.A.; Mechtler, L.; Gleason, M.J. Intravenous carboplatin for recurrent malignant glioma: A phase II study. J. Clin. Oncol. 1991, 9, 860–864. [Google Scholar] [CrossRef]
  40. Posadas, I.; Alonso-Moreno, C.; Bravo, I.; Carrillo-Hermosilla, F.; Garzon, A.; Villaseca, N.; Lopez-Solera, I.; Albaladejo, J.; Cena, V. Synthesis, Characterization, DNA Interactions and Antiproliferative Activity on Glioblastoma of Iminopyridine Platinum(II) Chelate Complexes. J. Inorg. Biochem. 2017, 168, 46–54. [Google Scholar] [CrossRef]
  41. Mishima, K.; Mazar, A.P.; Gown, A.; Skelly, M.; Ji, X.D.; Wang, X.D.; Jones, T.R.; Cavenee, W.K.; Huang, H.J. A peptide derived from the non-receptor binding region of urokinase plasminogen activator inhibits glioblastoma growth and angiogenesis in vivo in combination with cisplatin. Proc. Natl. Acad. Sci. USA 2000, 97, 8484–8489. [Google Scholar] [CrossRef] [PubMed]
  42. Spiro, T.P.; Liu, L.; Majka, S.; Haaga, J.; Willson, J.K.V.; Gerson, S.L. Temozolomide: The effect of once- and twice-a-day dosing on tumor tissue levels of the DNA repair protein O6-alkylguanine-DNA-alkyltransferase. Clin. Cancer Res. 2001, 7, 2309–2317. [Google Scholar] [PubMed]
  43. Zheng, Y.; Hunting, D.J.; Ayotte, P.; Sanche, L. Role of secondary lowenergy electrons in the concomitant chemoradiation therapy of cancer. Phys. Rev. Lett. 2008, 100, 198101. [Google Scholar] [CrossRef] [PubMed]
  44. Nagane, M.; Narita, Y.; Mishima, K.; Levitzki, A.; Burgess, A.W.; Cavenee, W.K.; Huang, H.J. Human Glioblastoma Xenografts Overexpressing a Tumor-Specific Mutant Epidermal Growth Factor Receptor Sensitized to Cisplatin by the Ag1478 Tyrosine Kinase Inhibitor. J. Neurosurg. 2001, 95, 472–479. [Google Scholar] [CrossRef]
  45. Brandes, A.A.; Basso, U.; Reni, M.; Vastola, F.; Tosoni, A.; Cavallo, G.; Scopece, L.; Ferreri, A.J.; Panucci, M.G.; Monfardini, S.; et al. First-Line Chemotherapy with Cisplatin Plus Fractionated Temozolomide in Recurrent Glioblastoma Multiforme: A Phase Ii Study of the Gruppo Italiano Cooperativo Di Neuro-Oncologia. J. Clin. Oncol. 2004, 22, 1598–1604. [Google Scholar] [CrossRef]
  46. Wang, X.; Wang, X.; Jin, S.; Muhammad, N.; Guo, Z. Stimuli-Responsive Therapeutic Metallodrugs. Chem. Rev. 2019, 119, 1138–1192. [Google Scholar] [CrossRef]
  47. Hess, S.M.; Anderson, J.G.; Bierbach, U. A Non-Crosslinking Platinum-Acridine Hybrid Agent Shows Enhanced Cytotoxicity Compared to Clinical Bcnu and Cisplatin in Glioblastoma Cells. Bioorg. Med. Chem. Lett. 2005, 15, 443–446. [Google Scholar] [CrossRef]
  48. Gajski, G.; Čimbora-Zovko, T.; Rak, S.; Osmak, M.; Garaj-Vrhovac, V. Antitumour Action on Human Glioblastoma A1235 Cells through Cooperation of Bee Venom and Cisplatin. Cytotechnology 2016, 68, 1197–1205. [Google Scholar] [CrossRef]
  49. Zhu, T.; Xu, Y.; Dong, B.; Zhang, J.; Wei, Z.; Xu, Y.; Yao, Y. Β-Elemene Inhibits Proliferation of Human Glioblastoma Cells through the Activation of Glia Maturation Factor Β and Induces Sensitization to Cisplatin. Oncol. Rep. 2011, 26, 405–413. [Google Scholar]
  50. Park, C.K.; Park, S.H.; Lee, S.H.; Kim, C.Y.; Kim, D.W.; Paek, S.H.; Kim, D.G.; Heo, D.S.; Kim, I.H.; Jung, H.W. Methylation Status of the Mgmt Gene Promoter Fails to Predict the Clinical Outcome of Glioblastoma Patients Treated with Acnu Plus Cisplatin. Neuropathology 2009, 29, 443–449. [Google Scholar] [CrossRef]
  51. Buckner, J.C.; Ballman, K.V.; Michalak, J.C.; Burton, G.V.; Cascino, T.L.; Schomberg, P.J.; Hawkins, R.B.; Scheithauer, B.W.; Sandler, H.M.; Marks, R.S.; et al. Phase Iii Trial of Carmustine and Cisplatin Compared with Carmustine Alone and Standard Radiation Therapy or Accelerated Radiation Therapy in Patients with Glioblastoma Multiforme: North Central Cancer Treatment Group 93-72-52 and Southwest Oncology Group 9503 Trials. J. Clin. Oncol. 2006, 24, 3871–3879. [Google Scholar] [PubMed]
  52. Barth, R.F.; Wu, G.; Hans Meisen, W.; Nakkula, R.J.; Yang, W.; Huo, T.; Kellough, D.A.; Kaumaya, P.; Turro, C.; Agius, L.M.; et al. Design, synthesis, and evaluation of cisplatincontaining EGFR targeting bioconjugates as potential therapeutic agents for brain tumors. OncoTargets Ther. 2016, 9, 2769–2781. [Google Scholar] [CrossRef] [PubMed]
  53. Wang, M.; Wu, Q.; Fang, M.; Huang, W.; Zhu, H. Mir-152-3p Sensitizes Glioblastoma Cells Towards Cisplatin Via Regulation of Sos1. OncoTargets Ther. 2019, 12, 9513–9525. [Google Scholar] [CrossRef] [PubMed]
  54. Li, J.; Song, J.; Guo, F. Mir-186 Reverses Cisplatin Resistance and Inhibits The formation of the Glioblastoma-Initiating Cell phenotype by Degrading Yin Yang 1 in Glioblastoma. Int. J. Mol. Med. 2019, 43, 517–524. [Google Scholar] [CrossRef] [PubMed]
  55. Baldwin, R.M.; Garratt-Lalonde, M.; Parolin, D.A.E.; Krzyzanowski, P.M.; Andrade, M.A.; Lorimer, I.A.J. Protection of Glioblastoma Cells from Cisplatin Cytotoxicity Via Protein Kinase Cι-Mediated Attenuation of P38 Map Kinase Signaling. Oncogene 2006, 25, 2909–2919. [Google Scholar] [CrossRef] [PubMed]
  56. Ding, L.; Yuan, C.; Wei, F.; Wang, G.; Zhang, J.; Bellail, A.C.; Zhang, Z.; Olson, J.J.; Hao, C. Cisplatin Restores Trail Apoptotic Pathway in Glioblastoma-Derived Stem Cells through up-Regulation of Dr5 and Down-Regulation of C-Flip. Cancer Investig. 2011, 29, 511–520. [Google Scholar] [CrossRef]
  57. Ferrari, B.; Urselli, F.; Gilodi, M.; Camuso, S.; Priori, E.C.; Rangone, B.; Ravera, M.; Veneroni, P.; Zanellato, I.; Roda, E.; et al. New Platinum-Based Prodrug Pt(Iv)Ac-Poa: Antitumour Effects in Rat C6 Glioblastoma Cells. Neurotox. Res. 2020, 37, 183–197. [Google Scholar] [CrossRef]
  58. Macieja, A.; Kopa, P.; Galita, G.; Pastwa, E.; Majsterek, I.; Poplawski, T. Comparison of the Effect of Three Different Topoisomerase Ii Inhibitors Combined with Cisplatin in Human Glioblastoma Cells Sensitized with Double Strand Break Repair Inhibitors. Mol. Biol. Rep. 2019, 46, 3625–3636. [Google Scholar] [CrossRef]
  59. Gwak, H.-S.; Shingu, T.; Chumbalkar, V.; Hwang, Y.-H.; DeJournett, R.; Latha, K.; Koul, D.; Alfred Yung, W.K.; Powis, G.; Farrell, N.P.; et al. Combined Action of the Dinuclear Platinum Compound Bbr3610 with the Pi3-K Inhibitor Px-866 in Glioblastoma. Int. J. Cancer 2011, 128, 787–796. [Google Scholar] [CrossRef]
  60. Rosa, P.; Catacuzzeno, L.; Sforna, L.; Mangino, G.; Carlomagno, S.; Mincione, G.; Petrozza, V.; Ragona, G.; Franciolini, F.; Calogero, A. Bk Channels Blockage Inhibits Hypoxia-Induced Migration and Chemoresistance to Cisplatin in Human Glioblastoma Cells. J. Cell. Physiol. 2018, 233, 6866–6877. [Google Scholar] [CrossRef]
  61. Benzina, S.; Fischer, B.; Miternique-Grosse, A.; Dufour, P.; Denis, J.M.; Bergerat, J.P.; Gueulette, J.; Bischoff, P. Cell Death Induced in a Human Glioblastoma Cell Line by P(65)+Be Neutrons Combined with Cisplatin. Life Sci. 2006, 79, 513–518. [Google Scholar] [CrossRef] [PubMed]
  62. Roci, E.; Cakani, B.; Brace, G.; Bushati, T.; Rroji, A.; Petrela, M.; Kaloshi, G. Platinum-based Chemotherapy in Recurrent High-grade Glioma Patients: Retrospective Study. Med. Arh. 2014, 68, 140–143. [Google Scholar] [CrossRef] [PubMed]
  63. Wang, Y.; Kong, X.; Guo, Y.; Wang, R.; Ma, W. Continuous Dose-Intense Temozolomide and Cisplatin in Recurrent Glioblastoma Patients. Medicine 2017, 96, e6261. [Google Scholar] [CrossRef] [PubMed]
  64. Wang, L.; Setlow, R.B. Inactivation of O-6-alkylguanine-DNA Alkyltransferase in HeLa cells by cisplatin. Carcinogenesis 1989, 10, 1681–1684. [Google Scholar] [CrossRef]
  65. Piccioni, D.; D’atri, S.; Papa, G.; Caravita, T.; Franchi, A.; Bonmassar, E.; Graziani, G. Cisplatin increases sensitivity of human leukemic blasts to triazene compounds. J. Chemother. 1995, 7, 224–229. [Google Scholar] [CrossRef]
  66. Britten, C.D.; Rowinsky, E.K.; Baker, S.D.; Agarwala, S.S.; Eckardt, J.R.; Barrington, R.; Diab, S.G.; Hammond, L.A.; Johnson, T.; Villalona-Calero, M.; et al. A phase I and pharmacokinetic study of Temozolomide and cisplatin in patients with advanced solid malignancies. Clin. Cancer Res. 1999, 5, 1629–1637. [Google Scholar]
  67. Rousseau, J.; Barth, R.F.; Fernandez, M.; Adam, J.-F.; Balosso, J.; Estève, F.; Elleaume, H. Efficacy of Intracerebral Delivery of Cisplatin in Combination with Photon Irradiation for Treatment of Brain Tumors. J. Neurooncol. 2010, 98, 287–295. [Google Scholar] [CrossRef]
  68. Billecke, C.; Finniss, S.; Tahash, L.; Miller, C.; Mikkelsen, T.; Farrell, N.P.; Bögler, O. Polynuclear platinum anticancer drugs are more potent than cisplatin and induce cell cycle arrest in glioma. Neuro Oncol. 2006, 8, 215–226. [Google Scholar] [CrossRef]
  69. Newton, H.B.; Page, M.A.; Junck, L.; Greenberg, H.S. Intra-arterial cisplatin for the treatment of malignant gliomas. J Neurooncol. 1989, 7, 39–45. [Google Scholar] [CrossRef]
  70. Feun, L.G.; Stewart, D.J.; Maor, M.; Leavens, M.; Savaraj, N.; Burgess, M.A.; Yung, W.K.A.; Benjamin, R.S. A pilot study of cis-diamminedichloroplatinum and radiation therapy in patients with high grade astrocytomas. J. Neurooncol. 1983, 1, 109–113. [Google Scholar] [CrossRef]
  71. Silvani, A.; Gaviani, P.; Lamperti, E.A.; Eoli, M.; Falcone, C.; DiMeco, F.; Milanesi, I.M.; Erbetta, A.; Boiardi, A.; Fariselli, L.; et al. Cisplatinum and BCNU chemotherapy in primary glioblastoma patients. J. Neurooncol. 2009, 94, 57–62. [Google Scholar] [CrossRef] [PubMed]
  72. Balaña, C.; López-Pousa, A.; Berrocal, A.; Yaya-Tur, R.; Herrero, A.; García, J.L.; Martín-Broto, J.; Benavides, M.; Cerdá-Nicolás, M.; Ballester, R.; et al. Phase II study of temozolomide and cisplatin as primary treatment prior to radiotherapy in newly diagnosed glioblastoma multiforme patients with measurable disease. A study of the Spanish Medical Neuro-Oncology Group (GENOM). J. Neurooncol. 2004, 70, 359–370. [Google Scholar] [CrossRef] [PubMed]
  73. Chovanec, M.; Abu Zaid, M.; Hanna, N.; El-Kouri, N.; Einhorn, L.H.; Albany, C. Long-term Toxicity of Cisplatin in Germ-Cell Tumor Survivors. Ann. Oncol. 2017, 28, 2670–2679. [Google Scholar] [CrossRef] [PubMed]
  74. Wang, D.; Wang, C.; Wang, L.; Chen, Y. A comprehensive review in improving delivery of small-molecule chemotherapeutic agents overcoming the blood-brain/brain tumor barriers for glioblastoma treatment. Drug Deliv. 2019, 26, 551–565. [Google Scholar] [CrossRef] [PubMed]
  75. Kovacs, Z.I.; Burks, S.R.; Frank, J.A. Focused ultrasound with microbubbles induces sterile inflammatory response proportional to the blood brain barrier opening: Attention to experimental conditions. Theranostics 2018, 8, 2245–2248. [Google Scholar] [CrossRef]
  76. Sharabi, S.; Bresler, Y.; Ravid, O.; Shemesh, C.; Atrakchi, D.; Schnaider-Beeri, M.; Gosselet, F.; Dehouck, L.; Last, D.; Guez, D.; et al. Transient blood–brain barrier disruption is induced by low pulsed electrical fields in vitro: An analysis of permeability and trans-endothelial electric resistivity. Drug Deliv. 2019, 26, 459–469. [Google Scholar] [CrossRef]
  77. Graham-Gurysh, E.; Moore, K.M.; Satterlee, A.B.; Sheets, K.T.; Lin, F.C.; Bachelder, E.M.; Miller, C.R.; Hingtgen, S.D.; Ainslie, K.M. Sustained Delivery of Doxorubicin Via Acetalated Dextran Scaffold Prevents Glioblastoma Recurrence after Surgical Resection. Mol. Pharm. 2018, 15, 1309–1318. [Google Scholar] [CrossRef]
  78. Tabet, A.; Jensen, M.P.; Parkins, C.C.; Patil, P.G.; Watts, C.; Scherman, O.A. Designing Next-Generation Local Drug Delivery Vehicles for Glioblastoma Adjuvant Chemotherapy: Lessons from the Clinic. Adv. Healthc. Mater. 2019, 8, e1801391. [Google Scholar] [CrossRef]
  79. Minghan, S.; Sanche, L. Convection-Enhanced Delivery in Malignant Gliomas: A Review of Toxicity and Efficacy. J. Oncol. 2019, 2019, 9342796. [Google Scholar]
  80. Mathew, E.N.; Berry, B.C.; Yang, H.W.; Carroll, R.S.; Johnson, M.D. Delivering Therapeutics to Glioblastoma: Overcoming Biological Constraints. Int. J. Mol. Sci. 2022, 23, 1711. [Google Scholar] [CrossRef]
  81. Ruiz-Molina, D.; Mao, X.; Alfonso-Triguero, P.; Lorenzo, J.; Bruna, J.; Yuste, V.J.; Candiota, A.P.; Novio, F. Advances in Preclinical/Clinical Glioblastoma Treatment: Can Nanoparticles Be of Help? Cancers 2022, 14, 4960. [Google Scholar] [CrossRef] [PubMed]
  82. Harder, B.G.; Blomquist, M.R.; Wang, J.; Kim, A.J.; Woodworth, G.F.; Winkles, J.A.; Loftus, J.C.; Tran, N.L. Developments in Blood-Brain Barrier Penetrance and Drug Repurposing for Improved Treatment of Glioblastoma. Front. Oncol. 2018, 8, 462. [Google Scholar] [CrossRef] [PubMed]
  83. Hersh, D.; Wadajkar, A.S.; Roberts, N.B.; Perez, J.G.; Connolly, N.P.; Frenkel, V.; Winkles, J.A.; Woodworth, G.F.; Kim, A.J. Evolving Drug Delivery Strategies to Overcome the Blood Brain Barrier. Curr. Pharm. Des. 2016, 22, 1177–1193. [Google Scholar] [CrossRef] [PubMed]
  84. McCarthy, D.J.; Malhotra, M.A.; O’Mahony, M.; Cryan, J.F.; O’Driscoll, C.M. Nanoparticles and the Blood-Brain Barrier: Advancing from In-Vitro Models Towards Therapeutic Significance. Pharm. Res. 2015, 32, 1161–1185. [Google Scholar] [CrossRef]
  85. Saraiva, C.; Praça, C.; Ferreira, R.; Santos, T.; Ferreira, L.; Bernardino, L. Nanoparticle-mediated brain drug delivery: Overcoming blood–brain barrier to treat neurodegenerative diseases. J. Control. Release 2016, 235, 34–47. [Google Scholar] [CrossRef]
  86. Zhang, Y.; Fu, X.; Jia, J.; Wikerholmen, T.; Xi, K.; Kong, Y.; Wang, J.; Chen, H.; Ma, Y.; Li, Z.; et al. Glioblastoma therapy using codelivery of cisplatin and glutathione peroxidase targeting siRNA from iron oxide nanoparticles. ACS Appl. Mater. Interf. 2020, 12, 43408–43421. [Google Scholar] [CrossRef]
  87. Wadajkar, A.S.; Dancy, J.G.; Hersha, D.S.; Anastasiadis, P.; Tran, N.L.; Woodworth, G.F.; Winkles, J.A.; Kim, A.J. Tumor-targeted Nanotherapeutics: Overcoming Treatment Barriers for Glioblastoma. Wiley Interdiscip. Rev. Nanomed. Nanobiotechnol. 2017, 9, e1439. [Google Scholar] [CrossRef]
  88. Wang, X.; Guo, Z. Targeting and delivery of platinum-based anticancer drugs. Chem. Soc. Rev. 2013, 42, 202–224. [Google Scholar] [CrossRef]
  89. Haxton, K.J.; Burt, H.M. Polymeric drug delivery of platinum-based anticancer agents. J. Pharm. Sci. 2009, 98, 2299–2316. [Google Scholar] [CrossRef]
  90. Wu, H.; Cabral, H.; Toh, K.; Mi, P.; Chen, Y.-C.; Matsumoto, Y.; Yamada, N.; Liu, X.; Kinoh, H.; Miura, Y. Polymeric micelles loaded with platinum anticancer drugs target preangiogenic micrometastatic niches associated with inflammation. J. Control Release 2014, 189, 1–10. [Google Scholar] [CrossRef]
  91. Kesavan, A.; Ilaiyaraja, P.; Beaula, W.S.; Kumari, V.V.; Lal, J.S.; Arunkumar, C.; Anjana, G.; Srinivas, S.; Ramesh, A.; Rayala, S.K.; et al. Tumor targeting using polyamidoamine dendrimer–cisplatin nanoparticles functionalized with diglycolamic acid and herceptin. Eur. J. Pharm. Biopharm. 2015, 96, 255–263. [Google Scholar] [CrossRef] [PubMed]
  92. Timbie, K.F.; Afzal, U.; Date, A.; Zhang, C.; Song, J.; Miller, G.W.; Suk, J.S.; Hanes, J.; Price, R.J. MR image-guided delivery of cisplatin-loaded brain-penetrating nanoparticles to invasive glioma with focused ultrasound. J. Control. Release 2017, 263, 120–131. [Google Scholar] [CrossRef]
  93. Thanasupawat, T.; Bergen, H.; Hombach-Klonisch, S.; Krcek, J.; Ghavami, S.; Del Bigio, M.R.; Krawitz, S.; Stelmack, G.; Halayko, A.; McDougall, M.; et al. Platinum (IV) coiled coil nanotubes selectively kill human glioblastoma cells. Nanomed. Nanotechnol. Biol. Med. 2015, 11, 913–925. [Google Scholar] [CrossRef] [PubMed]
  94. Ortiz-Islas, E.; Sosa-Arróniz, A.; Manríquez-Ramírez, M.E.; Rodríguez-Pérez, C.E.; Tzompantzi, F.; Padilla, J.M. Mesoporous silica nanoparticles functionalized with folic acid for targeted release Cis-Pt to glioblastoma cells. Rev. Adv. Mater. Sci. 2021, 60, 25–37. [Google Scholar] [CrossRef]
  95. Depciuch, J.; Miszczyk, J.; Maximenko, A.; Zielinski, P.M.; Rawojc, K.; Panek, A.; Olko, P.; Parlinska-Wojtan, M. Gold Nanopeanuts as Prospective Support for Cisplatin in Glioblastoma Nano-Chemo-Radiotherapy. Int. J. Mol. Sci. 2020, 21, 9082. [Google Scholar] [CrossRef] [PubMed]
  96. Mao, X.; Calero-Pérez, P.; Montpeyó, D.; Bruna, J.; Yuste, V.J.; Candiota, A.P.; Lorenzo, J.; Novio, F.; Ruiz-Molina, D. Intranasal Administration of Catechol-Based Pt(IV) Coordination Polymer Nanoparticles for Glioblastoma Therapy. Nanomaterials 2022, 12, 1221. [Google Scholar] [CrossRef] [PubMed]
  97. Charest, G.; Sanche, L.; Fortin, D.; Mathieu, D.; Paquette, B. Glioblastoma Treatment: Bypassing the Toxicity of Platinum Compounds by Using Liposomal Formulation and Increasing Treatment Efficiency with Concomitant, Radiotherapy. Int. J. Radiation Oncol. Biol. Phys. 2012, 84, 244–249. [Google Scholar] [CrossRef]
  98. Zahednezhad, F.; Zakeri-Milani, P.; Shahbazi Mojarrad, J.; Valizadeh, H. The latest advances of cisplatin liposomal formulations: Essentials for preparation and analysis. Expert Opin. Drug Deliv. 2020, 17, 523–541. [Google Scholar] [CrossRef]
  99. Liu, L.; Ye, Q.; Lu, M.; Lo, Y.-C.; Hsu, Y.-H.; Wei, M.-C.; Chen, Y.-H.; Lo, S.-C.; Wang, S.-J.; Bain, D.J. A new approach to reduce toxicities and to improve bioavailabilities of platinum-containing anti-cancer nanodrugs. Sci. Rep. 2015, 5, 10881. [Google Scholar] [CrossRef]
  100. Hernández-Pedro, N.Y.; Rangel-López, E.; Magaña-Maldonado, R.; Pérez de la Cruz, V.; Santamaría del Angel, A.; Pineda, B.; Sotelo, J. Application of Nanoparticles on Diagnosis and Therapy in Gliomas. Biomed. Res. Int. 2013, 2013, 351031. [Google Scholar] [CrossRef]
  101. Zhao, Y.Z.; Lin, Q.; Wong, H.L.; Shen, X.T.; Yang, W.; Xu, H.L.; Mao, K.L.; Tian, F.R.; Yang, J.J.; Xu, J.; et al. Glioma-targeted therapy using Cilengitide nanoparticles combined with UTMD enhanced delivery. J. Control. Release 2016, 224, 112–125. [Google Scholar] [CrossRef] [PubMed]
  102. Nainwal, N. Recent advances in transcranial focused ultrasound (FUS) triggered brain delivery. Curr. Drug Targets 2017, 18, 1225–1232. [Google Scholar] [CrossRef] [PubMed]
  103. Hoadley, K.A.; Yau, C.; Wolf, D.M.; Cherniack, A.D.; Tamborero, D.; Ng, S.; Leiserson, M.D.; Niu, B.; McLellan, M.D.; Uzunangelov, V. Multiplatform analysis of 12 cancer types reveals molecular classification within and across tissues of origin. Cell 2014, 158, 929–944. [Google Scholar] [CrossRef] [PubMed]
  104. Charest, G.; Sanche, L.; Fortin, D.; Mathieu, D.; Paquette, B. Optimization of the Route of Platinum Drugs Administration to Optimize the Concomitant Treatment with Radiotherapy for Glioblastoma Implanted in the Fischer Rat Brain. J. Neurooncol. 2013, 115, 365–373. [Google Scholar] [CrossRef]
  105. Rahman, A.A.; Stojanovska, V.; Pilowsky, P.; Nurgali, K. Platinum accumulation in the brain and alteration in the central regulation of cardiovascular and respiratory functions in oxaliplatin-treated rats. Pflugers Arch—Eur. J. Physiol. 2021, 473, 107–120. [Google Scholar] [CrossRef]
  106. Dermitzakis, E.V.; Kimiskidis, V.K.; Eleftheraki, A.; Lazaridis, G.; Konstantis, A.; Basdanis, G.; Tsiptsios, I.; Georgiadis, G.; Fountzilas, G. The impact of oxaliplatin-based chemotherapy for colorectal cancer on the autonomous nervous system. Eur. J. Neurol. 2014, 21, 1471–1477. [Google Scholar] [CrossRef]
  107. Tajbakhsh, M.; Houghton, P.J.; Morton, C.L.; Kolb, E.A.; Gorlick, R.; Maris, J.M.; Keir, S.T.; Wu, J.; Reynolds, C.P.; Smith, M.A.; et al. Initial testing of cisplatin by the pediatric preclinical testing program. Pediatr. Blood Cancer 2008, 50, 992–1000. [Google Scholar] [CrossRef]
  108. Rao, A.A.N.; Wallace, D.J.; Billups, C.; Boyett, J.M.; Gajjar, A.; Packer, R.J. Cumulative cisplatin dose is not associated with event-free or overall survival in children with newly diagnosed average-risk medulloblastoma treated with cisplatin based adjuvant chemotherapy: Report from the Children’s Oncology Group. Pediatr. Blood Cancer 2014, 61, 102–106. [Google Scholar] [CrossRef]
  109. Enríquez Pérez, J.; Fitzell, S.; Kopecky, J.; Visse, E.; Darabi, A.; Siesjö, P. The effect of locally delivered cisplatin is dependent on an intact immune function in an experimental glioma model. Sci. Rep. 2019, 9, 5632. [Google Scholar] [CrossRef]
  110. Xue, Y.; Gao, S.; Gou, J.; Yin, T.; He, H.; Wang, Y.; Zhang, Y.; Tang, X.; Wu, R. Platinum-based chemotherapy in combination with PD-1/PD-L1 inhibitors: Preclinical and clinical studies and mechanism of action. Expert Opin. Drug Deliv. 2021, 18, 187–203. [Google Scholar] [CrossRef]
  111. Park, S.J.; Ye, W.; Xiao, R.; Silvin, C.; Padget, M.; Hodge, J.W.; Van Waes, C.; Schmitt, N.C. Cisplatin and oxaliplatin induce similar immunogenic changes in preclinical models of head and neck cancer. Oral Oncol. 2019, 95, 127–135. [Google Scholar] [CrossRef]
  112. Eiseman, J.L.; Beumer, J.H.; Rigatti, L.H.; Strychor, S.; Meyers, K.; Dienel, S.; Horn, C.C. Plasma Pharmacokinetics and Tissue and Brain Distribution of Cisplatin in Musk Shrews. Cancer Chemother. Pharmacol. 2015, 75, 143–152. [Google Scholar] [CrossRef]
  113. McWhinney, S.R.; Goldberg, R.M.; McLeod, H.L. Platinum neurotoxicity pharmacogenetics. Mol. Cancer Ther. 2009, 8, 10–16. [Google Scholar] [CrossRef] [PubMed]
  114. Rademaker-Lakhai, J.M.; Crul, M.; Zuur, L.; Baas, P.; Beijnen, J.H.; Simis, Y.J.; Van Zandwijk, N.; Schellens, J.H. Relationship between cisplatin administration and the development of ototoxicity. J. Clin. Oncol. 2006, 24, 918–924. [Google Scholar] [CrossRef]
  115. Pardridge, W.M. The blood-brain barrier: Bottleneck in brain drug development. NeuroRx 2005, 2, 3–14. [Google Scholar] [CrossRef] [PubMed]
  116. Massimino, M.; Spreafico, F.; Riva, D.; Biassoni, V.; Poggi, G.; Solero, C.; Gandola, L.; Genitori, L.; Modena, P.; Simonetti, F.; et al. A lower-dose, lower-toxicity cisplatin-etoposide regimen for childhood progressive low-grade glioma. J. Neurooncol. 2010, 100, 65–71. [Google Scholar] [CrossRef] [PubMed]
  117. Ferrari, B.; Roda, E.; Priori, E.C.; De Luca, F.; Facoetti, A.; Ravera, M.; Brandalise, F.; Locatelli, C.A.; Rossi, P.; Bottone, M.G. A New Platinum-Based Prodrug Candidate for Chemotherapy and Its Synergistic Effect with Hadrontherapy: Novel Strategy to Treat Glioblastoma. Front. Neurosci. 2021, 15, 589906. [Google Scholar] [CrossRef]
  118. Prados, M.D.; Schold, S.C.; Fine, H.A.; Jaeckle, K.; Hochberg, F.; Mechtler, L.; Fetell, M.R.; Phuphanich, S.; Feun, L.; Janus, T.J.; et al. A randomized, double-blind, placebo-controlled, phase 2 study of RMP-7 in combination with carboplatin administered intravenously for the treatment of recurrent malignant glioma. Neuro Oncol. 2003, 5, 96–103. [Google Scholar] [CrossRef]
  119. Sheleg, S.; Korotkevich, E.; Zhavrid, E.; Muravskaya, G.; Smeyanovich, A.; Shanko, Y.; Yurkshtovich, T.; Bychkovsky, P.; Belyaev, S. Local chemotherapy with cisplatin-depot for glioblastoma multiforme. J. Neurooncol. 2002, 60, 53–59. [Google Scholar] [CrossRef]
  120. White, E.; Bienemann, A.; Taylor, H.; Hopkins, K.; Cameron, A.; Gill, S. A phase I trial of carboplatin administered by convection-enhanced delivery to patients with recurrent/progressive glioblastoma multiforme. Contemp. Clin. Trials 2012, 33, 320–331. [Google Scholar] [CrossRef]
  121. Saida, Y.; Watanabe, S.; Tanaka, T.; Baba, J.; Sato, K.; Shoji, S.; Igarashi, N.; Kondo, R.; Okajima, M.; Koshio, J.; et al. Critical roles of chemoresistant effector and regulatory T cells in antitumor immunity after lymphodepleting chemotherapy. J. Immunol. 2015, 195, 726–735. [Google Scholar] [CrossRef] [PubMed]
  122. Roberts, N.B.; Wadajkar, A.S.; Winkles, J.A.; Davila, E.; Kim, A.J.; Woodworth, G.F. Repurposing platinum-based chemotherapies for multi-modal treatment of glioblastoma. Oncoimmunology 2016, 5, e1208876. [Google Scholar] [CrossRef] [PubMed]
  123. Kreuter, J. Drug delivery to the central nervous system by polymeric nanoparticles: What do we know? Adv. Drug Deliv. Rev. 2014, 71, 2–14. [Google Scholar] [CrossRef] [PubMed]
  124. Johnstone, T.C.; Wilson, J.J.; Lippard, S.J. Monofunctional and Higher-Valent Platinum Anticancer Agents. Inorg. Chem. 2013, 52, 12234–12249. [Google Scholar] [CrossRef]
  125. Sun, H.; Chen, X.; Chen, D.; Dong, M.; Fu, X.; Li, Q.; Liu, X.; Wu, Q.; Qiu, T.; Wan, T.; et al. Influences of surface coatings and components of FePt nanoparticles on the suppression of glioma cell proliferation. Int. J. Nanomed. 2012, 7, 3295–3307. [Google Scholar]
  126. Liang, S.; Zhou, Q.; Wang, M.; Zhu, Y.; Wu, Q.; Yang, X. Water-soluble l-cysteine-coated FePt nanoparticles as dual MRI/CT imaging contrast agent for glioma. Int. J. Nanomed. 2015, 10, 2325–2333. [Google Scholar]
  127. Duan, X.; He, C.; Kron, S.J.; Lin, W. Nanoparticle formulations of cisplatin for cancer therapy. Wiley Interdiscip. Rev. NanomedNanobiotechnol. 2016, 8, 776–791. [Google Scholar] [CrossRef]
  128. Feldhaeusser, B.; Platt, S.R.; Marrache, S.; Kolishetti, N.; Pathak, R.K.; Montgomery, D.J.; Reno, L.R.; Howerth, E.; Dhar, S. Evaluation of nanoparticle delivered cisplatin in beagles. Nanoscale 2015, 7, 13822–13830. [Google Scholar] [CrossRef]
  129. Kim, J.-H.; Kim, Y.-S.; Park, K.; Lee, S.; Nam, H.Y.; Min, K.H.; Jo, H.G.; Park, J.H.; Choi, K.; Jeong, S.Y.; et al. Antitumor efficacy of cisplatin-loaded glycol chitosan nanoparticles in tumor-bearing mice. J. Control Release 2008, 127, 41–49. [Google Scholar] [CrossRef]
  130. Ding, D.; Zhu, Z.; Liu, Q.; Wang, J.; Hu, Y.; Jiang, X.; Liu, B. Cisplatin-loaded gelatin-poly(acrylic acid) nanoparticles: Synthesis, antitumor efficiency in vivo and penetration in tumors. Eur. J. Pharm. Biopharm. 2011, 79, 142–149. [Google Scholar] [CrossRef]
  131. Mattheolabakis, G.; Taoufik, E.; Haralambous, S.; Roberts, M.L.; Avgoustakis, K. In vivo investigation of tolerance and antitumor activity of cisplatin-loaded PLGA-mPEG nanoparticles. Eur. J. Pharm. Biopharm. 2009, 71, 190–195. [Google Scholar] [CrossRef] [PubMed]
  132. Song, W.; Li, M.; Tang, Z.; Li, Q.; Yang, Y.; Liu, H.; Duan, T.; Hong, H.; Chen, X. Methoxypoly(ethylene glycol)-block-poly(L-glutamic acid)-loaded cisplatin and a combination with iRGD for the treatment of non-small-cell lung cancers. Macromol. Biosci. 2012, 12, 1514–1523. [Google Scholar] [CrossRef] [PubMed]
  133. 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 cisplatinincorporated polymeric micelles (NC-6004) in patients with solid tumours. Br. J. Cancer 2011, 104, 593–598. [Google Scholar] [CrossRef] [PubMed]
  134. Nishiyama, N.; Yokoyama, M.; Aoyagi, T.; Okano, T.; Sakurai, Y.; Kataoka, K. Preparation and characterization of self-assembled polymer-metal complex micelle from cis-dichlorodiammineplatinum (II) and poly (ethylene glycol)-poly (α, β-aspartic acid) block copolymer in an aqueous medium. Langmuir 1999, 15, 377–383. [Google Scholar] [CrossRef]
  135. Ye, H.; Jin, L.; Hu, R.; Yi, Z.; Li, J.; Wu, Y.; Xi, X.; Wu, Z. Poly (γ, l-glutamic acid)–cisplatin conjugate effectively inhibits human breast tumor xenografted in nude mice. Biomaterials 2006, 27, 5958–5965. [Google Scholar] [CrossRef]
  136. Feng, Z.; Lai, Y.; Ye, H.; Huang, J.; Xi, X.G.; Wu, Z. Poly (γ, L-glutamic acid)-cisplatin bioconjugate exhibits potent antitumor activity with low toxicity: A comparative study with clinically used platinum derivatives. Cancer Sci. 2010, 101, 2476–2482. [Google Scholar] [CrossRef]
  137. Kirkpatrick, G.J.; Plumb, J.A.; Sutcliffe, O.B.; Flint, D.J.; Wheate, N.J. Evaluation of anionic half generation 3.5–6.5 poly(amidoamine) dendrimers as delivery vehicles for the active component of the anticancer drug cisplatin. J. Inorg. Biochem. 2011, 105, 1115–1122. [Google Scholar] [CrossRef]
  138. Yellepeddi, V.K.; Kumar, A.; Maher, D.M.; Chauhan, S.C.; Vangara, K.K.; Palakurthi, S. Biotinylated PAMAM dendrimers for intracellular delivery of cisplatin to ovarian cancer: Role of SMVT. Anticancer Res. 2011, 31, 897–906. [Google Scholar]
  139. Mallick, A.; More, P.; Ghosh, S.; Chippalkatti, R.; Chopade, B.A.; Lahiri, M.; Basu, S. Dual drug conjugated nanoparticle for simultaneous targeting of mitochondria and nucleus in cancer cells. ACS Appl. Mater. Interf. 2015, 7, 7584–7598. [Google Scholar] [CrossRef]
  140. Iinuma, H.; Maruyama, K.; Okinaga, K.; Sasaki, K.; Sekine, T.; Ishida, O.; Ogiwara, N.; Johkura, K.; Yonemura, Y. Intracellular targeting therapy of cisplatin-encapsulated transferrin-polyethylene glycol liposome on peritoneal dissemination of gastric cancer. Int. J. Cancer 2002, 99, 130–137. [Google Scholar] [CrossRef]
  141. Krieger, M.L.; Eckstein, N.; Schneider, V.; Koch, M.; Royer, H.-D.; Jaehde, U.; Bendas, G. Overcoming cisplatin resistance of ovarian cancer cells by targeted liposomes in vitro. Int. J. Pharm. 2010, 389, 10–17. [Google Scholar] [CrossRef] [PubMed]
  142. Burger, K.N.; Staffhorst, R.W.; de Vijlder, H.C.; Velinova, M.J.; Bomans, P.H.; Frederik, P.M.; de Kruijff, B. Nanocapsules: Lipid-coated aggregates of cisplatin with high cytotoxicity. Nat. Med. 2002, 8, 81–84. [Google Scholar] [CrossRef] [PubMed]
  143. Hamelers, I.H.; De Kroon, A.I. Nanocapsules: A novel lipid formulation platform for platinum-based anticancer drugs. J. Liposome Res. 2007, 17, 183–189. [Google Scholar] [CrossRef] [PubMed]
  144. Melanie, K.; Heike, Z.; Sibylle, B.S.; Hartmut, O.; Matthias, Z.; Christian, B.; Rudolf, H.; Karl-Jürgen, H.; Werner, A.K.; Ingrid, H. Characterization of iron oxide nanoparticles adsorbed with cisplatin for biomedical applications. Phys. Med. Biol. 2009, 54, 5109–5121. [Google Scholar]
  145. Comenge, J.; Romero, F.M.; Sotelo, C.; Domínguez, F.; Puntes, V. Exploring the binding of Pt drugs to gold nanoparticles for controlled passive release of cisplatin. J. Control. Release 2010, 148, e31–e32. [Google Scholar] [CrossRef]
  146. Craig, G.E.; Brown, S.D.; Lamprou, D.A.; Graham, D.; Wheate, N.J. Cisplatin-tethered gold nanoparticles that exhibit enhanced reproducibility, drug loading, and stability: A step closer to pharmaceutical approval? Inorg. Chem. 2012, 51, 3490–3497. [Google Scholar] [CrossRef]
  147. 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. Angew. Chem. Int. Ed. 2012, 51, 6742–6747. [Google Scholar] [CrossRef]
  148. Lin, C.-H.; Cheng, S.-H.; Liao, W.-N.; Wei, P.-R.; Sung, P.-J.; Weng, C.-F.; Lee, C.-H. Mesoporous silica nanoparticles for the improved anticancer efficacy of cisplatin. Int. J. Pharm. 2012, 429, 138–147. [Google Scholar] [CrossRef]
  149. 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. [Google Scholar] [CrossRef]
  150. Tao, Z.; Toms, B.; Goodisman, J.; Asefa, T. Mesoporous silica microparticles enhance the cytotoxicity of anticancer platinum drugs. ACS Nano 2010, 4, 789–794. [Google Scholar] [CrossRef]
  151. Vashist, S.K.; Zheng, D.; Pastorin, G.; Al-Rubeaan, K.; Luong, J.H.; Sheu, F.-S. Delivery of drugs and biomolecules using carbon nanotubes. Carbon 2011, 49, 4077–4097. [Google Scholar] [CrossRef]
  152. Li, J.; Yap, S.Q.; Yoong, S.L.; Nayak, T.R.; Chandra, G.W.; Ang, W.H.; Panczyk, T.; Ramaprabhu, S.; Vashist, S.K.; Sheu, F.S.; et al. Carbon nanotube bottles for incorporation, release and enhanced cytotoxic effect of cisplatin. Carbon 2012, 50, 1625–1634. [Google Scholar] [CrossRef]
  153. Guven, A.; Rusakova, I.A.; Lewis, M.T.; Wilson, L.J. Cisplatin@US-tube carbon nanocapsules for enhanced chemotherapeutic delivery. Biomaterials 2012, 33, 1455–1461. [Google Scholar] [CrossRef]
  154. Dhar, S.; Liu, Z.; Thomale, J.; Dai, H.; Lippard, S.J. Targeted single-wall carbon nanotube-mediated Pt(IV) prodrug delivery using folate as a homing device. J. Am. Chem. Soc. 2008, 130, 11467–11476. [Google Scholar] [CrossRef]
  155. Rieter, W.J.; Pott, K.M.; Taylor, K.M.L.; Lin, W. Nanoscale coordination polymers for platinum-based anticancer drug delivery. J. Am. Chem. Soc. 2008, 130, 11584–11585. [Google Scholar] [CrossRef]
  156. 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. [Google Scholar] [CrossRef]
  157. Adarsh, N.N.; Frias, C.; Ponnoth Lohidaksh, T.M.; Lorenzo, J.; Novio, F.; Garcia-Pardo, J.; Ruiz-Molina, D. Pt(IV)-based nanoscale coordination polymers: Antitumor activity, cellular uptake and interactions with nuclear DNA. Chem. Eng. J. 2018, 340, 94–102. [Google Scholar] [CrossRef]
  158. 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. [Google Scholar] [CrossRef]
  159. Makharza, S.A.; Cirillo, G.; Vittorio, O.; Valli, E.; Voli, F.; Farfalla, A.; Curcio, M.; Iemma, F.; Nicoletta, F.P.; El-Gendy, A.A.; et al. Magnetic Graphene Oxide Nanocarrier for Targeted Delivery of Cisplatin: A Perspective for Glioblastoma Treatment. Pharmaceuticals 2019, 12, 76. [Google Scholar] [CrossRef]
  160. Coluccia, D.; Figueiredo, C.A.; YiJun Wu, M.; Riemenschneider, A.N.; Diaz, R.; Luck, A.; Smith, C.; Das, S.; Ackerley, C.; O’Reilly, M.; et al. Enhancing glioblastoma treatment using cisplatin-gold-nanoparticle conjugates and targeted delivery with magnetic resonance-guided focused ultrasound. Nanomed. Nanotechnol. Biol. Med. 2018, 14, 1137–1148. [Google Scholar] [CrossRef]
  161. Gotov, O.; Battogtokh, G.; Shin, D.; Ko, J.T. Hyaluronic acid-coated cisplatin conjugated gold nanoparticles for combined cancer treatment. J. Ind. Eng. Chem. 2018, 65, 236–243. [Google Scholar] [CrossRef]
  162. Kosmas, C.; Angel, J.; Athanasiou, A.; Rapti, A.; Karanikas, C.; Lambaki, S.; Politis, N.; Mylonakis, N. 9088 Phase III study of Lipoplatin plus Gemcitabine versus Cisplatin plus Gemcitabine in advanced NSCLC; interim analysis. Eur. J. Cancer Suppl. 2009, 7, 531. [Google Scholar] [CrossRef]
  163. Farhat, F.; Kattan, J.; Ibrahim, K.; Bitar, N.; Haddad, N.; Tamraz, S.; Hatoum, H.; Shamseddine, A. 457 Preliminary results of a phase II study of lipoplatin (liposomal cisplatin)–vinorelbine combination as first line treatment in HER2/neu negative metastatic breast cancer (MBC). Eur. J. Cancer Suppl. 2010, 8, 192. [Google Scholar] [CrossRef]
  164. Koukourakis, M.I.; Giatromanolaki, A.; Pitiakoudis, M.; Kouklakis, G.; Tsoutsou, P.; Abatzoglou, I.; Panteliadou, M.; Sismanidou, K.; Sivridis, E.; Boulikas, T. Concurrent liposomal cisplatin (Lipoplatin), 5-fluorouracil and radiotherapy for the treatment of locally advanced gastric cancer: A phase I/II study. Int. J. Radiat. Oncol. Biol. Phys. 2010, 78, 150–155. [Google Scholar] [CrossRef]
  165. Huo, T.; Barth, R.F.; Yang, W.; Nakkula, R.J.; Koynova, R.; Tenchov, B.; Chaudhury, A.R.; Agius, L.; Boulikas, T.; Elleaume, H.; et al. Preparation, Biodistribution and Neurotoxicity of Liposomal Cisplatin following Convection Enhanced Delivery in Normal and F98 Glioma Bearing Rats. PLoS ONE 2012, 7, e48752. [Google Scholar] [CrossRef]
  166. Gonda, A.; Zhao, N.; Shah, J.V.; Calvelli, H.R.; Kantamneni, H.; Francis, N.L.; Ganapathy, V. Engineering Tumor-Targeting Nanoparticles as Vehicles for Precision Nanomedicine. Med. One 2019, 4, e190021. [Google Scholar]
  167. Shein, S.A.; Kuznetsov, I.I.; Abakumova, T.O.; Chelushkin, P.S.; Melnikov, P.A.; Korchagina, A.A.; Bychkov, D.A.; Seregina, I.F.; Bolshov, M.A.; Kabanov, A.V.; et al. VEGF- and VEGFR2-Targeted Liposomes for Cisplatin Delivery to Glioma Cells. Mol. Pharm. 2016, 13, 3712–3723. [Google Scholar] [CrossRef]
  168. Ashrafzadeh, M.S.; Akbarzadeh, A.; Heydarinasab, A.; Ardjmand, M. In vivo Glioblastoma Therapy Using Targeted Liposomal Cisplatin. Int. J. Nanomed. 2020, 15, 7035–7049. [Google Scholar] [CrossRef]
  169. Dhami, N.K.; Pandey, R.S.; Jain, U.K.; Chandra, R.; Madan, J. Non-Aggregated Protamine-Coated Poly(Lactide-Co-Glycolide) Nanoparticles of Cisplatin Crossed Blood-Brain Barrier, Enhanced Drug Delivery and Improved Therapeutic Index in Glioblastoma Cells: In Vitro Studies. J. Microencapsul. 2014, 31, 685–693. [Google Scholar] [CrossRef]
  170. Marrache, S.; Pathak, R.K.; Dhar, S. Detouring of cisplatin to access mitochondrial genome for overcoming resistance. Proc. Natl. Acad. Sci. USA 2014, 111, 10444–10449. [Google Scholar] [CrossRef]
  171. Nance, E.A.; Woodworth, G.F.; Sailor, K.A.; Shih, T.Y.; Xu, Q.; Swaminathan, G.; Xiang, D.; Eberhart, C.; Hanes, J. A dense poly(ethylene glycol) coating improves penetration of large polymeric nanoparticles within brain tissue. Sci. Transl. Med. 2012, 4, 149ra119. [Google Scholar] [CrossRef]
  172. Ebrahimi Shahmabadi, H.; Movahedi, F.; Koohi Moftakhari Esfahani, M.; Alavi, S.E.; Eslamifar, A.; Mohammadi Anaraki, G.; Akbarzadeh, A. Efficacy of Cisplatin-Loaded Polybutyl Cyanoacrylate Nanoparticles on the Glioblastoma. Tumour Biol. 2014, 35, 4799–4806. [Google Scholar] [CrossRef]
  173. Zhang, C.; Nance, E.A.; Mastorakos, P.; Chisholma, J.; Eberhart, S.B.C.; Tyler, B.; Brem, H.; Suk, J.S.; Hanes, J. Convection enhanced delivery of cisplatin-loaded brain penetrating nanoparticles cures malignant glioma in rats. J. Control. Release 2017, 263, 112–119. [Google Scholar] [CrossRef]
  174. Lajous, H.; Riva, R.; Lelievre, B.; Tetaud, C.; Avril, S.; Hindre, F.; Boury, F.; Jerome, C.; Lecomte, P.; Garcion, E. Hybrid Gd3+/Cisplatin Cross-Linked Polymer Nanoparticles Enhance Platinum Accumulation and Formation of DNA Adducts in Glioblastoma Cell Lines. Biomater. Sci. 2018, 6, 2386–2409. [Google Scholar] [CrossRef]
  175. Nukolova, N.V.; Baklaushev, V.P.; Abakumova, T.O.; Mel’nikov, P.A.; Abakumov, M.A.; Yusubalieva, G.M.; Bychkov, D.A.; Kabanov, A.V.; Chekhonin, V.P. Targeted delivery of cisplatin by connexin 43 vector nanogels to the focus of experimental glioma C6. Bull Exp. Biol. Med. 2014, 157, 524–529. [Google Scholar] [CrossRef]
  176. Renault-Mahieux, M.; Vieillard, V.; Seguin, J.; Espeau, P.; Le, D.T.; Lai-Kuen, R.; Mignet, R.; Paul, M.; Andrieux, K. Co-Encapsulation of Fisetin and Cisplatin into Liposomes for Glioma Therapy: From Formulation to Cell Evaluation. Pharmaceutics 2021, 13, 970. [Google Scholar] [CrossRef]
  177. Zou, Y.; Wang, Y.; Xu, S.; Liu, Y.; Yin, J.; Lovejoy, D.B.; Zheng, M.; Liang, X.-J.; Park, J.B.; Efremov, Y.M.; et al. Brain Co-Delivery of Temozolomide and Cisplatin for Combinatorial Glioblastoma Chemotherapy. Adv. Mater. 2022, 34, 2203958. [Google Scholar] [CrossRef]
  178. Straehla, J.P.; Hajal, C.; Safford, H.C.; Offeddu, G.S.; Boehnke, N.; Dacoba, T.C.; Wyckoff, J.; Kamm, R.D.; Hammond, P.T. A predictive microfluidic model of human glioblastoma to assess trafficking of blood-brain barrier penetrant nanoparticles. Proc. Natl. Acad. Sci. USA 2022, 119, e2118697119. [Google Scholar] [CrossRef]
  179. Yu, H.; Tang, Z.; Zhang, D.; Song, W.; Zhang, Y.; Yang, Y.; Ahmad, Z.; Chen, X. Pharmacokinetics, biodistribution and in vivo efficacy of cisplatin loaded poly(L-glutamic acid)-g-methoxy poly(ethylene glycol) complex nanoparticles for tumor therapy. J. Control Release 2015, 205, 89–97. [Google Scholar] [CrossRef]
  180. Ho, G.Y.; Woodward, N.; Coward, J.I.G. Cisplatin versus carboplatin: Comparative review of therapeuticmanagement in solid malignancies. Crit. Rev. Oncol./Hematol. 2016, 102, 37–46. [Google Scholar] [CrossRef]
  181. Colombo, N.; Guthrie, D.; Chiari, S.; Parmar, M.; Qian, W.; Swart, A.M.; Torri, V.; Williams, C.; Lissoni, A.; Bonazzi, C. International Collaborative Ovarian Neoplasm trial 1: A randomized trial of adjuvant chemotherapy in women with early-stage ovarian cancer. J. Natl. Cancer Inst. 2003, 95, 125–132. [Google Scholar] [PubMed]
  182. Rezaee, M.; Hunting, D.J.; Sanche, L. New insights into the mechanism underlying the synergistic action of ionizing radiation with platinum chemotherapeutic drugs: The role of low-energy electrons. Int. J. Radiat. Oncol. Biol. Phys. 2013, 87, 847–853. [Google Scholar] [CrossRef] [PubMed]
  183. Gianasi, E.; Buckley, R.G.; Latigo, J.; Wasil, M.; Duncan, R. HPMA copolymers platinates containing dicarboxylato ligands. Preparation, characterisation and in vitro and in vivo evaluation. J. Drug Target. 2002, 10, 549–556. [Google Scholar] [CrossRef] [PubMed]
  184. Gianasi, E.; Wasil, M.; Evagorou, E.; Keddle, A.; Wilson, G.; Duncan, R. HPMA copolymer platinates as novel antitumour agents: In vitro properties, pharmacokinetics and antitumour activity in vivo. Eur. J. Cancer. 1999, 35, 994–1002. [Google Scholar] [CrossRef]
  185. Rademaker-Lakhai, J.M.; Terret, C.; Howell, S.B.; Baud, C.M.; de Boer, R.F.; Pluim, D.; Beijnen, J.H.; Schellens, J.H.; 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. Clin. Cancer Res. 2004, 10, 3386–3395. [Google Scholar] [CrossRef]
  186. 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 Surf B Biointerfaces 2015, 130, 48–52. [Google Scholar] [CrossRef]
  187. Hassanzadeganroudsari, M.; Soltani, M.; Heydarinasab, A.; Apostolopoulos, V.; Akbarzadehkhiyavi, A.; Nurgali, K. Targeted nano-drug delivery system for glioblastoma therapy: In vitro and in vivo study. J. Drug Deliv. Sci. Technol. 2020, 60, 102039. [Google Scholar] [CrossRef]
  188. Arshad, A.; Yang, B.; Bienemann, A.S.; Barua, N.U.; Wyatt, M.J.; Woolley, M.; Johnson, D.E.; Edler, K.J.; Gill, S.S. Convection-enhanced delivery of carboplatin PLGA nanoparticles for the treatment of glioblastoma. PLoS ONE 2015, 10, e0132266. [Google Scholar] [CrossRef]
  189. Shi, M.; Anantha, M.; Wehbe, M.; Bally, M.B.; Fortin, D.; Roy, L.-O.; Charest, G.; Richer, M.; Paquette, B.; Sanche, L. Liposomal formulations of carboplatin injected by convection-enhanced delivery increases the median survival time of F98 glioma bearing rats. J. Nanobiotechnol. 2018, 16, 77. [Google Scholar] [CrossRef]
  190. Alex, A.T.; Joseph, A.; Shavi, G.; Rao, J.V.; Udupa, N. Development and evaluation of carboplatin-loaded PCL nanoparticles for intranasal delivery. Drug Deliv. 2016, 23, 2144–2153. [Google Scholar] [CrossRef]
  191. Fortin, D.; Desjardins, A.; Benko, A.; Niyonsega, T.; Boudrias, M. Enhanced chemotherapy delivery by intraarterial infusion and blood-brain barrier disruption in malignant brain tumors: The sherbrooke experience. Cancer 2005, 103, 2606–2615. [Google Scholar] [CrossRef] [PubMed]
  192. Miura, Y.; Takenaka, T.; Toh, K.; Wu, S.; Nishihara, H.; Kano, M.R.; Ino, Y.; Nomoto, T.; Matsumoto, Y.; Koyama, H.; et al. Cyclic Rgd-Linked Polymeric Micelles for Targeted Delivery of Platinum Anticancer Drugs to Glioblastoma through the Blood-Brain Tumor Barrier. ACS Nano 2013, 7, 8583–8592. [Google Scholar] [CrossRef]
  193. 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. Investig. New Drugs. 2016, 34, 269–276. [Google Scholar] [CrossRef]
  194. You, Y.; Wang, N.; He, l.; Shi, C.; Zhang, D.; Liu, Y.; Luo, L.; Che, T. Designing dual-functionalized carbon nanotubes with high blood–brain-barrier permeability for precise orthotopic glioma therapy. Dalton Trans. 2019, 48, 1569–1573. [Google Scholar] [CrossRef] [PubMed]
  195. McCartin, C.; Dussouillez, C.; Bernhard, C.; Mathieu, E.; Blumberger, J.; Dontenwill, M.; Herold-Mende, C.; Idbaih, A.; Lavalle, P.; Bellemin-Laponnaz, S.; et al. Polyethylenimine, an Autophagy-Inducing Platinum-Carbene-Based Drug Carrier with Potent Toxicity towards Glioblastoma Cancer Stem Cells. Cancers 2022, 14, 5057. [Google Scholar] [CrossRef] [PubMed]
  196. Wang, Z.; Denga, Z.; Zhu, G. Emerging platinum(IV) prodrugs to combat cisplatin resistance: From isolated cancer cells to tumor microenvironment. Dalton Trans. 2019, 48, 2536–2544. [Google Scholar] [CrossRef]
  197. Wang, Y.; Jiang, Y.; Wei, D.; Singh, P.; Yu, Y.; Lee, T.; Zhang, L.; Mandl, H.K.; Piotrowski-Daspit, A.S.; Chen, X.; et al. Nanoparticle-mediated convection-enhanced delivery of a DNA intercalator to gliomas circumvents temozolomide resistance. Nat. Biomed. Eng. 2021, 5, 1048–1058. [Google Scholar] [CrossRef]
  198. Arduino, I.; Depalo, N.; Re, F.; Dal Magro, R.; Panniello, A.; Margiotta, N.; Fanizza, E.; Lopalco, A.; Laquintana, V.; Cutrignelli, A.; et al. PEGylated solid lipid nanoparticles for brain delivery of lipophilic kiteplatin Pt(IV) prodrugs: An in vitro study. Int. J. Pharm. 2020, 583, 119351. [Google Scholar] [CrossRef]
  199. Mao, X.; Wu, S.; Calero-Pérez, P.; Candiota, A.P.; Alfonso, P.; Bruna, J.; Yuste, V.J.; Lorenzo, J.; Novio, F.; Ruiz-Molina, D. Synthesis and Validation of a Bioinspired Catechol-Functionalized Pt(IV) Prodrug for Preclinical Intranasal Glioblastoma Treatment. Cancers 2022, 14, 410. [Google Scholar] [CrossRef]
  200. Kutwin, M.; Sawosz, E.; Jaworski, S.; Hinzmann, M.; Wierzbicki, M.; Hotowy, A.; Grodzik, M.; Winnicka, A.; Chwalibog, A. Investigation of Platinum Nanoparticle Properties against U87 Glioblastoma Multiforme. Arch. Med. Sci. 2017, 13, 1322–1334. [Google Scholar] [CrossRef]
  201. Lopez Ruiz, A.; Bartomeu Garcia, C.; Navarro Gallon, S.; Webster, T.J. Novel Silver-Platinum Nanoparticles for Anticancer and Antimicrobial Applications. Int. J. Nanomed. 2020, 15, 169–179. [Google Scholar] [CrossRef]
  202. Shi, M.; Fortin, D.; Sanche, L.; Paquette, B. Convection-Enhancement Delivery of Platinum-Based Drugs and Lipoplatintm to Optimize the Concomitant Effect with Radiotherapy in F98 Glioma Rat Model. Investig. New Drugs 2015, 33, 555–563. [Google Scholar] [CrossRef] [PubMed]
  203. Setua, S.; Ouberai, M.; Piccirillo, S.G.; Watts, C.; Welland, M. Cisplatin-Tethered Gold Nanospheres for Multimodal Chemo-Radiotherapy of Glioblastoma. Nanoscale 2014, 6, 10865–10873. [Google Scholar] [CrossRef] [PubMed]
  204. Hall, P.E.; Lewis, R.; Syed, N.; Shaffer, R.; Evanson, J.; Ellis, S.; Williams, M.; Feng, X.; Johnston, A.; Thomson, J.A. A phase I study of Pegylated arginine deiminase (Pegargiminase), cisplatin, and Pemetrexed in Argininosuccinate Synthetase 1-deficient recurrent high-grade glioma. Clin. Cancer Res. 2019, 25, 2708–2716. [Google Scholar] [CrossRef]
  205. Allen, J.; Siffert, J.; Donahue, B.; Nirenberg, A.; Jakacki, R.; Robertson, P.; DaRosso, R.; Thoron, L.; Rosovsky, M.; Pinto, R. A phase I/II study of carboplatin combined with hyperfractionated radiotherapy for brainstem gliomas. Cancer Interdiscip. Int. J. Am. Cancer Soc. 1999, 86, 1064–1069. [Google Scholar] [CrossRef]
  206. Fouladi, M.; Blaney, S.M.; Poussaint, T.Y.; Freeman III, B.B.; McLendon, R.; Fuller, C.; Adesina, A.M.; Hancock, M.L.; Danks, M.K.; Stewart, C. Phase II study of oxaliplatin in children with recurrent or refractory medulloblastoma, supratentorial primitive neuroectodermal tumors, and atypical teratoid rhabdoid tumors: A pediatric brain tumor consortium study. Cancer Interdiscip. Int. J. Am. Cancer Soc. 2006, 107, 2291–2297. [Google Scholar] [CrossRef]
  207. Macy, M.E.; Duncan, T.; Whitlock, J.; Hunger, S.P.; Boklan, J.; Narendran, A.; Herzog, C.; Arceci, R.J.; Bagatell, R.; Trippett, T. A multi-center phase Ib study of oxaliplatin (NSC# 266046) in combination with fluorouracil and leucovorin in pediatric patients with advanced solid tumors. Pediatr. Blood Cancer 2013, 60, 230–236. [Google Scholar] [PubMed]
  208. Chintala, S.; Ali-Osman, F.; Mohanam, S.; Rayford, A.; Go, Y.; Gokaslan, Z.L.; Gagercas, E.; Vekaiah, B.; Sawaya, R.; Nicolson, G.; et al. Effect of cisplatin and BCNU on MMP-2 levels in human glioblastoma cell lines in vitro. Clin. Exp. Metastasis 1997, 14, 361–367. [Google Scholar] [CrossRef] [PubMed]
  209. Rébé, C.; Demontoux, L.; Pilot, T.; Ghiringhelli, F. Platinum Derivatives Effects on Anticancer Immune Response. Biomolecules 2020, 10, 13. [Google Scholar] [CrossRef]
  210. Nduom, E.K.; Weller, M.; Heimberger, A.B. Immunosuppressive mechanisms in glioblastoma. Neuro Oncol. 2015, 17, vii9–vii14. [Google Scholar] [CrossRef]
  211. Hato, S.V.; Khong, A.; de Vries, I.J.; Lesterhuis, W.J. Molecular pathways: The immunogenic effects of platinum-based chemotherapeutics. Clin. Cancer Res. 2014, 20, 2831–2837. [Google Scholar] [CrossRef] [PubMed]
  212. Lesterhuis, W.J.; Punt, C.J.; Hato, S.V.; Eleveld-Trancikova, D.; Jansen, B.J.; Nierkens, S.; Schreibelt, G.; de Boer, A.; Van Herpen, C.M.; Kaanders, J.H.; et al. Platinum-based drugs disrupt STAT6-mediated suppression of immune responses against cancer in humans and mice. J. Clin. Investig. 2011, 121, 3100–3108. [Google Scholar] [CrossRef]
  213. Bezu, L.; Gomes-de-Silva, L.C.; Dewitte, H.; Breckpot, K.; Fucikova, J.; Spisek, R.; Galluzzi, L.; Kepp, O.; Kroemer, G. Combinatorial strategies for the induction of immunogenic cell death. Front. Immunol. 2015, 6, 187. [Google Scholar] [CrossRef]
  214. Kang, L.; Gao, Z.; Huang, W.; Jin, M.; Wang, Q. Nanocarrier-mediated codelivery of chemotherapeutic drugs and gene agents for cancer treatment. Acta Pharm. Sin. B 2015, 5, 169–175. [Google Scholar] [CrossRef] [PubMed]
  215. Field, K.M.; Simes, J.; Nowak, A.K.; Cher, L.; Wheeler, H.; Hovey, E.J.; Brown, C.S.; Barnes, E.H.; Sawkins, K.; Livingstone, A.; et al. Randomized phase 2 study of carboplatin and bevacizumab in recurrent glioblastoma. Neuro Oncol. 2015, 17, 1504–1513. [Google Scholar] [CrossRef] [PubMed]
  216. Bruinsmann, F.A.; Richter Vaz, G.; de Cristo Soares Alves, A.; Aguirre, T.; Raffin Pohlmann, A.; Stanisçuaski Guterres, S.; Sonvico, F. Nasal Drug Delivery of Anticancer Drugs for the Treatment of Glioblastoma: Preclinical and Clinical Trials. Molecules 2019, 24, 4312. [Google Scholar] [CrossRef]
  217. Sandbhor, P.; Goda, J.; Mohanty, B.; Gera, P.; Yadav, S.; Chekuri, G.; Chaudhari, P.; Dutt, S.; Banerjee, R. Targeted nano-delivery of chemotherapy via intranasal route suppresses in vivo glioblastoma growth and prolongs survival in the intracranial mouse model. Drug Deliv. Transl. 2023, 13, 608–626. [Google Scholar] [CrossRef]
Figure 1. Structured of Clinically Approved Platinum Anticancer Drugs. Reproduced from Ref. [37] with permission of the copyright holder.
Figure 1. Structured of Clinically Approved Platinum Anticancer Drugs. Reproduced from Ref. [37] with permission of the copyright holder.
Nanomaterials 13 01619 g001
Figure 2. Deactivation and/or sequestration pathways of cisplatin. In extracellular environment, through interaction with serum albumin; once it is inside the cell, sequestration by coordination to metal-thioneins (MT) or inactivation through reaction with glutathione (GSH). This inactivation generates adducts removed from the cell through specific pumps such as ATP7B, implicated in cisplatin resistance. Reproduced from Ref. [124] with permission of the copyright holder.
Figure 2. Deactivation and/or sequestration pathways of cisplatin. In extracellular environment, through interaction with serum albumin; once it is inside the cell, sequestration by coordination to metal-thioneins (MT) or inactivation through reaction with glutathione (GSH). This inactivation generates adducts removed from the cell through specific pumps such as ATP7B, implicated in cisplatin resistance. Reproduced from Ref. [124] with permission of the copyright holder.
Nanomaterials 13 01619 g002
Figure 3. Preparation scheme of HA-CisPt-AuNp. AuNp: gold nanoparticle; CisPt: cisplatin or cis-diamminedichloroplatinum(II); DPA: 3,30-dithiodipropionic acid; HA: hyaluronic acid; and PEG: polyethylene glycol. Reproduced from Ref. [161] with permission of the copyright holder.
Figure 3. Preparation scheme of HA-CisPt-AuNp. AuNp: gold nanoparticle; CisPt: cisplatin or cis-diamminedichloroplatinum(II); DPA: 3,30-dithiodipropionic acid; HA: hyaluronic acid; and PEG: polyethylene glycol. Reproduced from Ref. [161] with permission of the copyright holder.
Nanomaterials 13 01619 g003
Figure 4. Schematic representation of targeted PBCA NPs conjugated with monoclonal antibodies against epidermal growth factor receptors (EGFR) and the platinum-based complexes drug delivery process. Reproduced from Ref. [187] with permission of the copyright holder.
Figure 4. Schematic representation of targeted PBCA NPs conjugated with monoclonal antibodies against epidermal growth factor receptors (EGFR) and the platinum-based complexes drug delivery process. Reproduced from Ref. [187] with permission of the copyright holder.
Nanomaterials 13 01619 g004
Figure 5. (a) Schematic showing the design strategy for the oxaliplatin encapsulation on the PEGl-glutamic acid (Glu) micelles functionalized with cyclic Arg-Gly-Asp (cRGD), and proposed transvascular transport for the NPs; (b) In vitro cytotoxicity, in U87 cells, of free oxaliplatin or resulting micelles with different percentages of targeting molecules (the data were analyzed using Student’s t-test, and * p < 0.001 was considered significant; NS: not significant); (c) Representative in vivo bioluminescence images of mice treated with 20% non-targeted and 20% of targeted micelles. Reproduced from Ref. [192] with permission of the copyright holder.
Figure 5. (a) Schematic showing the design strategy for the oxaliplatin encapsulation on the PEGl-glutamic acid (Glu) micelles functionalized with cyclic Arg-Gly-Asp (cRGD), and proposed transvascular transport for the NPs; (b) In vitro cytotoxicity, in U87 cells, of free oxaliplatin or resulting micelles with different percentages of targeting molecules (the data were analyzed using Student’s t-test, and * p < 0.001 was considered significant; NS: not significant); (c) Representative in vivo bioluminescence images of mice treated with 20% non-targeted and 20% of targeted micelles. Reproduced from Ref. [192] with permission of the copyright holder.
Nanomaterials 13 01619 g005
Figure 6. In vitro efficacy of T-Platin-M-NPs and comparison with cisplatin, carboplatin, Platin-M, and NT-Platin-M-NPs in two canine brain tumour cell lines, J3TBG and SDT3G. Cell viability was assessed by MTT assay after treatment for either 24 h followed by 48 h incubation or treatment for 72 h, depending on the drug concentration. The data are mean ± SD from three to five independent experiments. *** p < 0.001; ** p < 0.001–0.01. Reproduced from Ref. [128] with permission of the copyright holder.
Figure 6. In vitro efficacy of T-Platin-M-NPs and comparison with cisplatin, carboplatin, Platin-M, and NT-Platin-M-NPs in two canine brain tumour cell lines, J3TBG and SDT3G. Cell viability was assessed by MTT assay after treatment for either 24 h followed by 48 h incubation or treatment for 72 h, depending on the drug concentration. The data are mean ± SD from three to five independent experiments. *** p < 0.001; ** p < 0.001–0.01. Reproduced from Ref. [128] with permission of the copyright holder.
Nanomaterials 13 01619 g006
Figure 7. (a) Scheme of the Pt-Fe NCPs synthesis upon polymerization of a Pt(IV) prodrug with iron ions as metal nodes; Cumulative release profiles of Pt from Pt-Fe NCPs at 37 °C at (b) pH 7.4 and (c) pH 5.5 in PBS using the dialysis method in the absence or in the presence of glutathione (GSH); (d) Biodistribution of Pt-Fe NCPs in mice organs 1 h after administration (dosage of 1.5 mg/kg, n = 3); (e) Tumour volume evolution in the period 0–20 days post-implantation (p.i.) [blue: control; red: therapy starting point at day 10 p.i.; pink: therapy starting point at day 6 p.i.; *: case of cured mice with therapy starting point at day 6 p.i. Reproduced from Ref. [96] with permission of the copyright holder.
Figure 7. (a) Scheme of the Pt-Fe NCPs synthesis upon polymerization of a Pt(IV) prodrug with iron ions as metal nodes; Cumulative release profiles of Pt from Pt-Fe NCPs at 37 °C at (b) pH 7.4 and (c) pH 5.5 in PBS using the dialysis method in the absence or in the presence of glutathione (GSH); (d) Biodistribution of Pt-Fe NCPs in mice organs 1 h after administration (dosage of 1.5 mg/kg, n = 3); (e) Tumour volume evolution in the period 0–20 days post-implantation (p.i.) [blue: control; red: therapy starting point at day 10 p.i.; pink: therapy starting point at day 6 p.i.; *: case of cured mice with therapy starting point at day 6 p.i. Reproduced from Ref. [96] with permission of the copyright holder.
Nanomaterials 13 01619 g007
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Alfonso-Triguero, P.; Lorenzo, J.; Candiota, A.P.; Arús, C.; Ruiz-Molina, D.; Novio, F. Platinum-Based Nanoformulations for Glioblastoma Treatment: The Resurgence of Platinum Drugs? Nanomaterials 2023, 13, 1619. https://0-doi-org.brum.beds.ac.uk/10.3390/nano13101619

AMA Style

Alfonso-Triguero P, Lorenzo J, Candiota AP, Arús C, Ruiz-Molina D, Novio F. Platinum-Based Nanoformulations for Glioblastoma Treatment: The Resurgence of Platinum Drugs? Nanomaterials. 2023; 13(10):1619. https://0-doi-org.brum.beds.ac.uk/10.3390/nano13101619

Chicago/Turabian Style

Alfonso-Triguero, Paula, Julia Lorenzo, Ana Paula Candiota, Carles Arús, Daniel Ruiz-Molina, and Fernando Novio. 2023. "Platinum-Based Nanoformulations for Glioblastoma Treatment: The Resurgence of Platinum Drugs?" Nanomaterials 13, no. 10: 1619. https://0-doi-org.brum.beds.ac.uk/10.3390/nano13101619

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop