Next Article in Journal
Supported Nanostructured MoxC Materials for the Catalytic Reduction of CO2 through the Reverse Water Gas Shift Reaction
Next Article in Special Issue
Facile One-Pot Green Synthesis of Magneto-Luminescent Bimetallic Nanocomposites with Potential as Dual Imaging Agent
Previous Article in Journal
The Poynting Vector Field Generic Singularities in Resonant Scattering of Plane Linearly Polarized Electromagnetic Waves by Subwavelength Particles
Previous Article in Special Issue
Evaluation of Dosimetric Effect of Bone Scatter on Nanoparticle-Enhanced Orthovoltage Radiotherapy: A Monte Carlo Phantom Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Metal-Polymer Nanoconjugates Application in Cancer Imaging and Therapy

by
André Q. Figueiredo
1,
Carolina F. Rodrigues
1,
Natanael Fernandes
1,
Duarte de Melo-Diogo
1,
Ilídio J. Correia
1,* and
André F. Moreira
1,2,*
1
CICS-UBI—Health Sciences Research Centre, Universidade da Beira Interior, Av. Infante D. Henrique, 6200-506 Covilhã, Portugal
2
CPIRN-UDI/IPG—Centro de Potencial e Inovação em Recursos Naturais, Unidade de Investigação para o Desenvolvimento do Interior do Instituto Politécnico da Guarda, Avenida Dr. Francisco de Sá Carneiro, No. 50, 6300-559 Guarda, Portugal
*
Authors to whom correspondence should be addressed.
Nanomaterials 2022, 12(18), 3166; https://doi.org/10.3390/nano12183166
Submission received: 26 August 2022 / Revised: 8 September 2022 / Accepted: 9 September 2022 / Published: 13 September 2022

Abstract

:
Metallic-based nanoparticles present a unique set of physicochemical properties that support their application in different fields, such as electronics, medical diagnostics, and therapeutics. Particularly, in cancer therapy, the plasmonic resonance, magnetic behavior, X-ray attenuation, and radical oxygen species generation capacity displayed by metallic nanoparticles make them highly promising theragnostic solutions. Nevertheless, metallic-based nanoparticles are often associated with some toxicological issues, lack of colloidal stability, and establishment of off-target interactions. Therefore, researchers have been exploiting the combination of metallic nanoparticles with other materials, inorganic (e.g., silica) and/or organic (e.g., polymers). In terms of biological performance, metal-polymer conjugation can be advantageous for improving biocompatibility, colloidal stability, and tumor specificity. In this review, the application of metallic-polymer nanoconjugates/nanohybrids as a multifunctional all-in-one solution for cancer therapy will be summarized, focusing on the physicochemical properties that make metallic nanomaterials capable of acting as imaging and/or therapeutic agents. Then, an overview of the main advantages of metal-polymer conjugation as well as the most common structural arrangements will be provided. Moreover, the application of metallic-polymer nanoconjugates/nanohybrids made of gold, iron, copper, and other metals in cancer therapy will be discussed, in addition to an outlook of the current solution in clinical trials.

Graphical Abstract

1. Introduction

The manipulation of matter at the nanoscale led to the development of various nanomaterials, based on polymers, lipids, ceramics, or metals, that exhibit exciting properties for improving the performance of electronics, medical diagnostics, and therapeutics [1,2]. Particularly, the application of nanoparticles for cancer therapy led to the development of new alternatives from therapeutics (i.e., drug delivery, photothermal and photodynamic effects) to diagnostic and imaging, leveraging the innate ability of these materials to accumulate in the tumor tissue [3,4,5].
In this field, the application of metallic-based nanomaterials (e.g., gold, iron, silver, and copper) has been capturing more attention in recent years due to their characteristic physical (e.g., magnetic behavior, plasmonic resonance, and imaging capacity) and chemical (e.g., radical oxygen species (ROS) generation, and catalytic activity enhancement) properties, which render them as platforms suitable for creating multifunctional cancer therapeutics [2,6,7,8,9]. Moreover, there are several methodologies described for the synthesis of metallic nanoparticles, such as physical or chemical vapor deposition, sol-gel methods, chemical reduction, hydrothermal methods, solvothermal method, laser ablation, and green synthesis processes, allowing the selection of the process most compatible with the available laboratory/industrial conditions and desired physicochemical properties [10,11,12,13]. Additionally, metallic nanomaterials surface functionalization, a strategy often used to refine the therapeutics pharmacokinetics, can be performed by well-known and defined methodologies [14]. Thus, the metallic nanomaterials’ theranostic potential provides an all-in-one solution for cancer diagnosis, therapy, and real-time monitoring, which ultimately can improve the therapeutic outcome of anticancer therapy [15,16]. This is the rationale behind various metallic-based nanomaterials under clinical trials, such as Aurolase®, Nanotherm®, and Magnablate®.
Nevertheless, despite the wide number of publications showing the appealing features of metallic-based nanomaterials, their translation into the clinic is still very limited [17]. Such is widely associated with some toxicological issues, lack of colloidal stability, and the establishment of off-target interactions [18]. Additionally, when subjected to high-energy radiations, the metallic nanoparticles often undergo reshaping processes or are even degraded leading to the loss of their therapeutic potential [19]. Therefore, researchers have been exploring the combination of metallic nanoparticles with other materials, inorganic (e.g., silica) and/or organic (e.g., polymers). Particularly, the combination of the metallic nanoparticles’ physicochemical properties with the superior biological performance of synthetic or natural polymers emerges as a valuable and straightforward approach to develop more effective anti-cancer therapeutics [16,20]. In this review, the application of metallic-polymer nanoconjugates/nanohybrids as a multifunctional all-in-one solution for cancer therapy will be summarized. Initially, the physicochemical properties rendering the metallic nanomaterials’ potential to act as imaging and/or therapeutic agents will be summarized. Then, an overview of the main advantages of metal-polymer conjugation as well as the most common structural arrangements will be provided. Moreover, the application of metallic-polymer nanoconjugates/nanohybrids made of gold, iron, copper, and others in cancer therapy will be discussed, in addition to an outlook of the current solution in clinical trials.

2. Metallic Nanoparticles Applications and Therapies

Nanosized metals present optical and electrical properties that differentiate them from other nanomaterials, supporting their application in biomedicine, such as the development of biosensors (e.g., diagnosis of viruses using colloidal gold nanoparticles), bioimaging agents (e.g., iron oxide-based contrast agents), catalysts, mechanical reinforcement, and drug delivery system or other therapeutics. Moreover, these unique characteristics can be explored to create more effective antitumoral nanomedicines. The high density and X-ray attenuation capacity of metallic nanoparticles allow them the intrinsic capacity to be applied as contrast agents for bioimaging applications [21]. Up to now, several studies available in the literature have already shown that metallic nanoparticles provide a higher contrast enhancement in X-ray computed tomography (CT) imaging than the iodine-based contrast agents conventionally used in the clinic [22,23,24,25,26]. On the other hand, nanosized metals, such as gold, silver, and copper, show a pronounced plasmonic resonance phenomenon, i.e., the collective oscillation of the conduction band electrons in metal-based nanomaterials in response to the incident photons [27]. This interaction can lead to light absorption or scattering and is dependent on the size, morphology, distance, and dielectric constant of the metallic nanoparticles and surrounding medium [28,29,30]. In turn, the excited surface electrons can decay to the ground state via different processes (e.g., electron-to-photon, electron-to-electron, and electron-to-phonon energy conversion), the two most common events being the release of the absorbed energy in the form of light or heat [31]. The former is often explored to enhance the quantum efficiency and photostability of fluorophores, allowing the detection of lower quantities of biomarkers used in biosensing or bioimaging [32]. The latter is the foundation stone for the application of metallic nanoparticles in cancer hyperthermia/photothermal therapy [33]. However, it is essential to tailor the nanomaterials to interact specifically with near-infrared (NIR) radiation, a region of the spectra where the major biological components (e.g., collagen, hemoglobin, and water) have the lowest or insignificant absorption [34]. This will reduce the off-target interactions and guarantee a site-specific activation of the metallic nanomaterials. Then, the heat generated by the light-nanoparticles interaction can mediate the destruction of the cancer cells [35,36]. The elevation of the tumor temperature to values superior to 45 °C provokes irreversible damage to cancer cells (e.g., DNA degradation, cell membrane disruption, protein denaturation), leading to cell death (i.e., tumor ablation). Otherwise, if mild temperature increases are achieved (i.e., between 40 and 45 °C), the cell damage is less pronounced and often reversed by the cell repair mechanisms [5,37,38]. Nevertheless, this creates a time window during which the cancer cells are more sensitive to the action of other therapeutic modalities such as chemotherapy [39]. Furthermore, metallic nanomaterials, such as those composed of iron, nickel, and cobalt, can also present magnetic properties, also allowing their application as contrast agents (magnetic resonance imaging (MRI)) and in tumor magnetic hyperthermia [40]. This capacity to be magnetically manipulated by external magnetic fields is also explored to guide these metallic nanomaterials in the human body and promote a tumor-specific accumulation [41,42]. At the tumor site, the utilization of alternating magnetic fields will promote the nanoparticles’ vibration and consequently a localized temperature increase will be obtained [43].
Metallic nanomaterials have also shown the capacity to mediate the formation of ROS [18]. This oxidative stress can influence several cellular processes/structures, e.g., intracellular calcium concentrations, activate transcription factors, induce DNA damage, and lipid peroxidation (cell membrane disruption), and increased amounts of ROS are highly cytotoxic [18]. The mechanism of ROS generation by metallic nanomaterials is influenced by their physicochemical properties (e.g., size, chemical structure, surface area, and charge) [44]. Generally, metallic nanomaterials act as the reactant or catalyst for the reduction of molecular oxygen to water, which yields the production of ROS, such as superoxide radicals and hydroxyl radicals [45]. The ROS generation of metallic nanomaterials can be further boosted by light absorption [46]. During this process, the electrons transit to higher energy bands, facilitating the reaction with water or molecular oxygen and consequently the ROS generation, a process denominated by photodynamic effect [47,48].
Despite the imaging and therapeutic potential of metallic nanoparticles, the in vivo application and translation to the clinic are severely hindered by their low colloidal stability, high reactivity, the formation of the protein corona, and high cytotoxicity [49,50,51]. Therefore, to overcome these limitations, researchers have been combining the superior physicochemical features of metallic nanoparticles with the increased biological properties (e.g., biocompatibility, enhanced blood circulation time, targeting capacity) of synthetic and natural polymers, often referred to as metal-polymer nanocomplexes or nanohybrids.

3. Metal-Polymer Based Nanomaterials

The metal-polymer nanocomplexes or nanohybrids are a class of nanomaterials that aim to address the biological bottlenecks of the metallic nanoparticles’ administration in humans. For that purpose, several strategies have been explored to create these new nanomaterials (Figure 1), namely the (i) surface coating of metallic cores (e.g., physical linkage and layer-by-layer), (ii) entrapment of metallic cores within polymeric matrices (e.g., nanoparticle in situ growth or post-synthesis entrapment), and (iii) the utilization of polymeric capsules [52,53,54]. One of the main rationales behind the introduction of polymers is to increase the colloidal stability of metallic nanoparticles and avoid protein adsorption during the nanoparticles’ circulation in the blood [55,56].
Highly hydrophilic polymers are often associated with improvements in the nanoparticles’ blood circulation time, namely by minimizing the nanoparticle-protein interaction, avoiding nanoparticles’ aggregation as well as the size and charge changes, and reducing the recognition by the immune cells [57,58]. For this purpose, polymers, such as poly(ethylene glycol) (PEG), poly(oxazolines) (POX), and poly(zwitterion)s, have been excelling in enhancing the nanoparticles’ blood circulation [59]. The PEG and POX anti-fouling properties, which are attributed to two main events, (i) the steric repulsion and (ii) the water barrier, impede the nanoparticles–proteins interaction, avoiding the formation of a protein corona that negatively impacts the nanomaterials’ biological performance [60,61]. In turn, the zwitterionic polymers present overall electrostatic neutrality and high chain hydration that confer to them a stealthing capacity [62,63]. The higher blood circulation times associated with the utilization of these polymers will increase the nanomaterials’ probability to accumulate and interact with the tumor cells, which can be essential for achieving superior antitumoral effects [64]. On the other hand, the natural and synthetic polymers can also imprint a stimuli-responsive (e.g., pH, temperature, and enzymatic) behaviour on the metallic-based nanomaterials, which can be particularly advantageous for controlling drug delivery [33,65,66,67,68]. In this regard, the heat generated by the nanomaterials (e.g., PTT and magnetic hyperthermia) can be also explored to induce phase changes in the polymers (e.g., Poly(N-isopropylacrylamide) (PNIPAM)) or increase their solubility, which will trigger the drug release [68,69]. Additionally, at the tumor site, the metal-polymer nanocomplexes/nanohybrids will also have an impact on the physiological conditions that can be explored to trigger the drug release, such as an acidic pH, overexpression of certain enzymes (e.g., matrix metalloproteinases), and an increased RedOx environment [65,67,70]. Apart from the passive accumulation of the nanomaterials at the tumor site, usually dependent on the enhanced permeability and retention (EPR) effect, the polymers or other targeting moieties can confer a specific recognition of molecules overexpressed in the tumor tissue [5,71]. This higher specificity towards cancer cells will favor the accumulation of nanomaterials in these areas as well as the nanomaterials–cancer cell interaction [72]. Therefore, the metal–polymer nanocomplexes or nanohybrids show the potential to create novel and more effective anticancer therapeutics. In the following section, the application of metallic–polymer nanoconjugates/nanohybrids made of gold, iron, copper, and others in cancer therapy will be overviewed, showing the therapeutic modalities that each metal nanoparticle allows to explore.

3.1. Gold-Polymer Conjugates

Gold nanoparticles are one of the most explored metallic nanomaterials for biomedical applications, from bioimaging to drug delivery [73,74,75]. The applicability of gold nanomaterials in bioimaging is demonstrated in different works in the literature, principally as a contrast agent for CT imaging [76]. For example, Xi and co-workers observed that, when an X-ray beam of 100 KeV is used, gold nanoparticles present an absorption coefficient two-times superior to that obtained with iodine (a conventional contrast agent used in the clinic) [77]. In turn, the pronounced surface plasmon resonance of gold nanoparticles has been supporting the development of novel photothermal therapies for cancer therapy (Table 1). This phenomenon can be fine-tuned to enhance the gold nanoparticles’ absorption in the NIR region (700–1100 nm) by optimizing the nanoparticles’ size and/or shape (e.g., spheres, cages, stars, and rods) [37,78]. Per se, the application of gold nanospheres in cancer photothermal therapy (PTT) is limited since its absorption peak is in the 500–550 nm region [79]. However, the organization of gold nanospheres in nanoclusters or shells renders a shift in the absorption spectra to the NIR region [2,5]. This change in the gold nanoparticles’ optical properties is attributed to the coupling of the plasmon resonance of adjacent gold nanospheres [80]. Therefore, the fine-tuning of the gold nanoclusters or shells’ absorption spectra can be achieved by optimizing the nanospheres’ size and interparticle distance [2]. In turn, the surface plasmon resonance in non-spherical gold nanoparticles varies with the nanoparticle surface [81,82]. For example, the two different surfaces in gold nanorods, longitudinal and transversal surfaces, originate two absorption bands. The transversal surface leads to an absorption band in the 500–550 nm region of the spectra, whereas the longitudinal surface originates an absorption peak that can be fine-tuned from the visible to the NIR region of the spectra [83,84]. This peak generated by the longitudinal surface resonance is determined by the aspect ratio of gold nanorods (i.e., quotient between length and width) [85]. Otherwise, the surface plasmon resonance of gold nanostars is defined by the particle’s core size, the number of tips, as well as the tips’ length and width [2]. Therefore, the gold nanoparticles’ plasticity and fine-tunning ability to present a high absorption efficiency in the NIR region of the spectra propelled their application in the cancer PTT. Nevertheless, despite the theragnostic potential of gold nanoparticles, their direct application in the human body is hindered by their high affinity to establish interactions with thiol groups, which favors the interaction with different biomolecules and lead to the nanoparticles’ aggregation [80]. Moreover, the gold nanoparticles can also be degraded when exposed to high-energy radiations, such as those used in CT imaging and PTT, causing the loss of their bioactivity [86]. Furthermore, several reports in the literature also describe that gold nanoparticles are strongly accumulated in the kidneys, causing nephrotoxicity, and may also trigger the lysis of red blood cells [78]. To address these issues, researchers have been exploring the development of gold-polymer nanocomplexes or nanohybrids to increase colloidal stability, biocompatibility, and even tumor specificity.
Peng and colleagues demonstrated the applicability of gold-based nanomaterials in bioimaging by following the biodistribution and tumor accumulation of PEGylated dendrimer-entrapped gold nanospheres [91]. In fact, the authors reported that these nanomaterials have an attenuation intensity higher than Omnipaque (an iodine-based contrast agent), which allowed them to follow the PEGylated dendrimer-entrapped gold nanoparticles in the blood circulation, after intravenous injection, as well as to perform the CT imaging of SPC-A1 xenograft tumors. Moreover, the authors reported that the PEGylated dendrimer-entrapped gold nanoparticles had a half-decay time in the blood circulation of 31.76 h, which was 2.5 times higher than that of bare gold nanorods previously reported. Furthermore, Gu et al. produced RGD-modified mPEG-PLGA nanocapsules containing gold nanoclusters and indocyanine green for the imaging and PTT of breast cancer [92]. The loaded mPEG-PLGA nanocapsules were formed via a water-in-oil-in-water emulsion using sonication, where (i) the first emulsion consisted of the water phase with gold nanoclusters and the oil phase with indocyanine green and the mPEG-PLGA. In the second emulsion, the mPEG-PLGA nanocapsules were further modified with poly(vinyl alcohol) and poly(acrylic acid), allowing the subsequent functionalization with RGD peptide via carbodiimide chemistry. The authors demonstrated that both one-photon and two-photon imaging techniques could be used to follow the nanocapsules in 4T1 tumor-bearing BALB/c mice. Moreover, the RGD-modified nanocapsules showed a preferential accumulation in U87-MG cancer cells (overexpressing ανβ3 integrins), when compared to MCF-7 cancer cells (low expression of ανβ3 integrins), leading to the almost complete ablation of the cancer cells after irradiation with a NIR laser (808 nm, 2 W cm−2, for 5 min). Feng and co-workers developed two different tumor-targeted gold nanocages for the combinatorial chemo-PTT of breast cancer [68]. For that purpose, pH-responsive gold nanocages were formulated via electrostatic interaction between the poly(acrylic acid) and the surface of the particles, entrapping the gold particles in the polymeric chains (pAAu nanoparticles) and loaded with Erlotinib (Erl), an epidermal growth factor receptor (EGFR) inhibitor. In turn, temperature-responsive gold nanocages were produced by reacting them with the thiol-terminated N-isopropylacrylamide (NIPAM) and acrylamide (AM) (p(NIPAM-co-AM) co-polymer (pNAu nanoparticles) and subsequently loaded with doxorubicin (Dox). The pAAu nanoparticles showed a pH-triggered Erl release due to the protonation of poly(acrylic acid) in acidic pH (i.e., loosening of the polymeric barrier), 4.5%, 24.8%, 44.1%, or 66.3% Erl released after 6 h at pH 7.4, 6.5, 6, or 5. Otherwise, the Dox-loaded pNAu nanoparticles were responsive to the irradiation with a NIR laser (808 nm, 0.5 W cm−2, for 10 min) and consequent increase in temperature (i.e., superior to 45 °C, a value higher than the lower critical solution temperature). The authors reported 46.2% of Dox released in 10 h, after a 10 min NIR laser irradiation, contrasting with the 5% detected in the absence of NIR irradiation. The in vivo studies performed in MCF-7 and A431 tumor-bearing mice demonstrated a passive and preferential accumulation in the tumor tissue (Figure 2). Moreover, the combinatorial therapy led to the reduction of A431 tumors’ size by 98%, after 14 days, whereas in MCF-7 tumors (low expression of EGFR), these nanomaterials only slowed the tumor growth.

3.2. Iron-Polymer Conjugates

The utilization of iron oxide nanoparticles can be explored in different biomedical applications, such as drug delivery, hyperthermia, and magnetic resonance imaging. Similar to gold nanoparticles, iron oxide can be produced in different shapes, such as spherical, rod-like, and cubical [93,94]. These nanoparticles (Fe2O3 and Fe3O4), when their size is inferior to 20 nm, present a superparamagnetic behavior at room temperature, i.e., the magnetization of the nanoparticles is close to 0 in the absence of an external magnetic field [95,96]. The iron oxide nanoparticles magnetism allows for widespread application in cancer therapy (Table 2), namely as contrast agents for magnetic resonance imaging: Feridex®, Resovist®, and Endorem®. The data available in the literature indicate that iron oxide nanoparticles are less toxic than the conventionally used gadolinium-based contrast agents, without presenting significant losses in imaging capacity [95]. Moreover, the magnetic properties of iron oxide nanoparticles have also been explored to direct the accumulation of the nanoparticles, specifically towards the tumor tissue, and in certain cases mediate a hyperthermic effect [97,98]. The former explores the application of an external magnetic field to guide the nanoparticles and promote their accumulation in the target tissue [99]. The latter employs an alternating magnetic field to induce the oscillation of iron oxide nanoparticles, which in turn generate heat [97]. This hyperthermic effect is dependent on the magnetic properties of the nanoparticles as well as on the frequency and intensity of the alternating magnetic field [100,101]. However, the biological application of iron oxide nanoparticles is hindered by their limited colloidal stability, tending to agglomerate when in contact with biological fluids [102]. With this in mind, Xu et al. prepared GSH-responsive hyaluronic acid-coated small iron oxide nanoparticles (HIONPs) for the diagnosis of liver metastases [103]. The nanoparticles were formed via a one-pot method, promoting the oxidation of ferrous ions to create iron oxide nanoparticles that were coated with hyaluronic acid modified with dopamine through the establishment phenol−metal coordination interactions. The in vitro measurements in a 0.52 T NMR instrument showed that the HIONPs have a longitudinal proton relaxivity (r1) of 41.3 Fe mM−1 s−1, indicating the applicability of HIONPs as T1 MRI contrast agents. This was confirmed using a 3 T MR scanner where the HIONPs led to higher T1-weighted magnetic resonance signals and lower T2-weighted magnetic resonance signals when the Fe concentration increased from 0 to 0.2 mM. Moreover, the application of HIONPs (Fe—0.03 mmol kg−1) in mice with B16F10, 4T1, and CT26 liver metastases allowed the metastases detection via MRI, showing the highest contrast-to-noise ratio 1 h after injection. This capacity is attributed to the higher GSH concentration in the mice liver tissue (i.e., 11 to 76-times higher) when compared to the tumor/metastases. The higher GSH concentration promotes the removal of the hyaluronic acid coating and consequent nanoparticle aggregation, which led to a significant decrease in the r1 value. In this way, the hepatic tissues became dark whereas the tumor/metastases are bright in the MRI. Moreover, the authors also described that the HIONPs presented a higher imaging capacity than Primovist® (Gd-based imaging agent), where the metastases and surrounding liver tissue presented a similar contrast-to-noise ratio. In another work, Xiao and co-workers developed ultrathin vesicles with multimodal imaging capacity for the combinatorial chemo-PTT of cancer [104]. For that purpose, ultrasmall superparamagnetic iron oxide nanoparticles, cisplatin, and liquid perfluorohexane were encapsulated in PLGA nanovesicles, followed by the formation of an ultrathin silica layer to prevent leakage. Then, the surface of the particles was modified with polyaniline (a photothermal agent) and functionalized with R8-RGD. In these nanoparticles, the iron oxide acted as a T2-weighted magnetic resonance contrast agent, with a transverse relaxivity (r2) value of 258.5 Fe mM−1 s−1, three times superior to the iron oxide nanoparticles alone. Moreover, the in vivo studies in A549 tumor-bearing mice showed that the administration of the PLGA vesicles containing the iron oxide nanoparticles allows the monitoring of the changes in the tumor cellularity via MRI. Additionally, the chemo-photothermal combinatorial treatment led to a significant regression, close to 97%, of the tumor volume in 21 days. Chen and colleagues produced a dual-targeted magnetic iron oxide nanoparticle for the imaging and hyperthermia of breast tumors (Figure 3) [105]. The iron oxide nanoparticles were coated with a DSPE-PEG2000 shell and then modified with RGDyK (neovascular endothelium targeting) and D-glucosamine (glucose transporter affinity) via carbodiimide chemistry. The authors observed that the combination of magnetic and active targeting approaches resulted in the best contrast effect on T2-weighted MRI from 3 to 48 h, showing the tumor region completely dark due to the accumulation of the nanoparticles. The tumor/normal tissue signal ratio at 48 h for active targeting strategies was 0.6 whereas for the magnetic plus active targeting combination this value was inferior to 0.3, showing a higher difference in the signal obtained in the tumor and normal tissues. Moreover, the utilization of alternating current magnetic fields (1.485 × 109 Am−1 s−1), focused on the tumor region, induced a temperature increase to 44 °C, which successfully slowed the growth of 4T1 tumors when compared to the control group (i.e., the relative tumor volume of 500% and 200% for control and iron oxide groups, respectively).

3.3. Copper-Polymer Conjugates

Copper nanomaterials have emerged in recent years as promising inorganic nanoparticles for biomedical applications (Table 3). Copper is a transition metal and can be engineered to form various nanomaterials, such as copper oxides, copper selenides, and copper sulfides [111,112,113]. Among them, copper sulfides have been the most explored due to the simple synthesis process and high NIR absorbing capacity, allowing their application in cancer PTT [37]. On the other hand, the copper nanomaterials can also be used as Fenton-like reagents mediating the formation of ROS (chemodynamic therapy) [114]. Moreover, the generation of ROS can also be stimulated under light irradiation (photodynamic therapy (PDT)) [115]. However, copper is often defined as more toxic than iron and gold, which makes the release of copper ions in the human body undesirable [116,117]. With that in mind, Li et al. showed that the surface functionalization with an amphiphilic polymer, poly(isobutylene-alt-maleic anhydride) (PMA), enhances the colloidal stability of copper telluride nanoparticles without visible agglomeration for periods longer than one month [118]. Furthermore, in vitro studies performed in 3T3 fibroblasts showed significant cytotoxicity after irradiation with a NIR laser (830 nm, 0.5 mW cm−2, for 2 s). In another work, Li and colleagues developed a PEGylated copper sulfide nanoparticle for the simultaneous PDT and PTT of lung cancer (Figure 4) [119]. For that purpose, thiolated-PEG was reacted with the copper sulfide nanoparticles rendering the PEGylated nanoparticles. The authors observed that the nanoparticles’ irradiation (30 µg mL−1) with a NIR laser (808 nm, 1 W cm−2, for 10 min) reaches temperatures superior to 42 °C. Moreover, the authors also observed the continuous quenching of the p-nitrosodimethylaniline (RNO) absorption under NIR irradiation, indicating the generation of ROS during this period. Moreover, the in vivo studies in SPC-A-1 tumor-bearing mice showed that the combination of PDT and PTT decreases tumor growth, observing a 5% increase in the tumor volume at day 14 after administering PEGylated copper sulfide nanoparticles (30 µg mL−1) plus NIR. Similarly, Shi et al. produced a PEGylated copper sulfide nanoparticle modified with RGD to target metastatic gastric cancer [120]. The nanoparticles were formed by the reaction of thiolated-PEG-COOH and copper sulfide, followed by the RGD modification using carbodiimide chemistry. The obtained nanoparticles showed computed-tomography contrast capacity similar to the Iodixanol (a clinically used contrast agent). Furthermore, the irradiation of the nanoparticles (60 µg mL−1) with a NIR laser (808 nm, 1 W cm−2, for 5 min) led to a temperature increase to 60 °C. In the in vivo studies, the authors observed that the RGD-modified PEGylated copper sulfide nanoparticles allowed the identification of MKN45 tumors and metastasis in sentinel lymph nodes through T2-weighted magnetic resonance images. Moreover, the NIR laser irradiation (808 nm, 1 W cm−2, for 10 min) promoted the increase in tumor/metastases temperature to 57 °C, leading to the complete ablation of the metastatic MKN45 tumors (sentinel lymph nodes weight similar to the healthy ones, 2.5 mg).

3.4. Other Metal-Polymer Nanoconjugates

Apart from the previously presented gold-, iron-, and copper-based nanomaterials, other metals have also been explored to create novel and more effective anticancer therapeutics (Table 4). Zinc-based nanoparticles, particularly zinc oxide (ZnO), are considered relatively biocompatible and generally regarded as safe. These nanoparticles present photoluminescence properties and a band gap that facilitates the interaction with oxygen and hydroxyl ions prompting the generation of superoxide and hydroxyl radicals (Table 4) [130,131]. Moreover, this ROS generation shows a certain selectivity towards cancer cells, decreasing the potential for inducing side effects [132]. Song and co-workers functionalized ZnO nanoparticles with polyvinylpyrrolidone for application in the imaging and therapy of colon cancer [133]. These authors reported that the surface functionalization maintained the stability of nanoparticles for 14 days, whereas non-coated ZnO nanoparticles aggregated after three days, without impacting the ROS generation capacity. Moreover, the authors also observed that the administration of the polyvinylpyrrolidone coated ZnO nanoparticles, at a concentration of 50 µg mL−1, reduced the viability of SW480 cancer cells to 54% due to ROS generation. In turn, the ROS generation was boosted by the irradiation with UV light, with cell viability of 15% at a concentration of 50 µg mL−1 (IC50 of 21.688 µg mL−1). This PDT capacity was also observed in the SW480 tumor-bearing mice, where the nanoparticles plus light treatment slowed the tumor growth for 28 days, a tumor inhibition rate of 61.1%.
Platinum (Pt) is a catalytic noble metal that has been explored in cancer therapy, namely in the form of chemotherapeutic drugs containing platinum atoms, such as cisplatin and derivatives [134]. The utilization of platinum nanoparticles is focused on the release of platinum ions that will induce DNA damage and provoke cell death [135]. Additionally, the platinum nanoparticles can also mediate the generation of ROS or act as photothermal agents [136,137]. Chen and colleagues developed PEGylated porous platinum nanoparticles loaded with Dox for application in breast cancer therapy [138]. In the synthesis process, Pluronic F127 was used as a surfactant for the platinum nanoparticles and then reacted with thiolated-mPEG. This surface modification enhanced the solubility and stability of the platinum nanoparticles. Moreover, the authors also reported that the presence of a 10 mHz square wave AC field (10 mA) further enhanced the ROS generation capacity. In turn, the combination of Dox delivery and ROS generation induced the regression of 4T1 tumors, showing a tumor growth inhibition of 95.5% after 14 days. Zhu and co-workers prepared sodium hyaluronate stabilized platinum nanoparticles (HA/Pt) for mediating a photothermal effect [139]. Apart from the nanoparticle stabilization, the HA functionalization increased the nanoparticles’ specificity towards the MDA-MB231 cancer cells (overexpression of CD44), when compared to the uptake by NIH3T3 cells (low expression of CD44). Furthermore, the in vivo studies in MDA-MB231 tumor-bearing mice revealed that upon irradiation with a NIR laser (808 nm, 1 W cm−1, for 10 min), the HA/Pt nanoparticles induced an increase in the temperature of the tumor to 44 °C, which translated to a reduction in the tumor growth for 14 days.
Silver (Ag) is a noble metal with vast applications, being widely applied in the biomedical field as an antimicrobial agent [140]. The silver nanoparticles can mediate ROS formation and consequently induce lipid peroxidation, protein oxidation, and DNA damage [141,142]. Park and co-workers developed indocyanine green-loaded silver nanoparticles functionalized with PEG and BSA for application in cancer PTT [143]. The BSA was used to stabilize the produced silver nanoparticles, followed by the reaction with NHS-PEG for obtaining the functionalized silver nanoparticles. The authors reported that the PEG-BSA silver nanoparticles were stable in solution for at least five days after the synthesis. Moreover, the combination of indocyanine green-silver nanoparticles resulted in a more stable photothermal effect, reaching temperatures of ≈45 °C even after three irradiations with a NIR laser (885 nm, 1.3 W, for 10 min). The in vivo studies performed in B16F10 tumor-bearing mice showed a preferential accumulation in the liver, kidney, and tumors and upon irradiation (885 nm, 0.95 W for 20 min), the tumors’ temperature reached 49 °C. This photothermal effect led to the tumors’ ablation after four days.

3.5. Clinical Trials

Since the 1990s, more than 50 nanomedicines have been approved by the Food and Drug Administration (FDA) and are currently on the market [151]. However, in the last decade, only a small number of formulations successfully reached the clinic for cancer treatment, a fact that is in part related to the poor pharmacokinetics of the nanomaterials [152]. The latest data indicate that less than 1% of the administered nanoparticles reach the tumor site [153]. Nevertheless, most of the approved nanomedicines are based on liposomes, protein nanoparticles, nano-emulsions, and metal oxide nanoparticles.
Particularly, iron oxide nanoparticles’ clinical utility is demonstrated by the FDA approval for application in cancer diagnosis, hyperthermia, and iron deficiency anemia. Among them, it is worth highlighting the various contrast agents based on iron oxide nanoparticles commercially available for MRI, such as Feridex®, Resovist®, and Endorem®. Moreover, there are more systems under clinical trial, such as Magnablate® and Nanotherm® (Table 5). The former is an iron oxide nanoparticle developed for magneto-hyperthermia applications that underwent a Phase 0 clinical trial (ClinicalTrials.gov Identifier: NCT02033447), with 12 participants, for the thermoablation of prostate cancer (no results are yet available) [154]. In turn, Nanotherm® is also based on an aqueous suspension of iron oxide nanoparticles and was already approved by the European Medicines Agency (EMA) for brain tumor treatment [37]. Moreover, recently, a new clinical trial (ClinicalTrials.gov Identifier: NCT05010759; still recruiting) was announced to study the application of Nanotherm® in the ablation of prostate carcinoma.
Regarding gold-based nanoparticles, to this date, two promising nanomedicines are currently in clinical trials [151]. AuroLase® uses the particles denominated as AuroShell®, a PEGylated silica-gold nanoshell with ≈150 nm in size, for the laser-activated thermal ablation of solid tumors, metastatic lung tumors, and cancer prostatic tissue [151,155]. In the clinical trial (ClinicalTrials.gov Identifier: NCT01679470), a single dose of AuroShell® was administered to promote the ablation of primary and/or metastatic lung tumors, still, to this date, no results were posted [156]. There is a second clinical trial (ClinicalTrials.gov Identifier: NCT00848042) involving the utilization of this nanomedicine for the treatment of patients with refractory and/or recurrent head and neck tumors. The patients were subjected to one or more doses of laser irradiation (808 nm) [30,33]. The participants were divided into three groups: (i) five participants, Auroshell® dose 4.5 mL kg−1, laser potency 3.5 W; (ii) five participants, Auroshell® dose 7.5 mL kg−1, laser potency 4.5 W; (iii) one participant, Auroshell® dose 7.5 mL kg−1, laser potency 5 W. However, no data were published to assess the effect on the targeted tumors. In a more recent clinical trial (ClinicalTrials.gov Identifier: NCT02680535), the Auroshell® nanoparticles were tested in combination with the MRI/Ultrasound fusion technology to promote the focal ablation of neoplastic prostate tissue. The extension of this clinical trial is still active (ClinicalTrials.gov Identifier: NCT04240639), however no data have been found describing the evolution of the targeted tumors. The NU-0129® is another gold-based nanomedicine under clinical trial (ClinicalTrials.gov Identifier: NCT03020017). This nanoparticle is formed by a spherical gold nanoparticle conjugated with siRNA oligonucleotides for targeting BCL2L12 oncogene in glioblastoma multiforme or gliosarcoma treatment applications [151,157]. The early results showed that the NU-0129® nanoparticles can cross the blood–brain barrier without unexpected adverse effects, still pending the data regarding antitumoral efficacy [158].

4. Conclusions

Metallic-based nanomaterials have been showing promising results in a variety of cancer-related applications, from imaging to the ablation of tumors. Nevertheless, several of these nanomedicines remain in a preclinical stage, an exception being the application of iron oxide-derived nanomaterials in the imaging of tumors.
In this review, the unique set of physicochemical properties that make the metallic nanoparticles highly promising for biomedical applications is described. The high density and X-ray attenuation capacity make these nanomaterials natural contrast agents for conventional imaging techniques such as CT. Moreover, the surface plasmon resonance and/or the magnetism allow the utilization of hyperthermia treatments (e.g., PTT and magneto-hyperthermia). Therefore, the conjugation with natural and/or synthetic polymers can further increase the biological performance of the metallic nanoparticles by enhancing the colloidal stability and blood circulation time as well as conferring additional specificity to the cancer cells. Nevertheless, there remain significant challenges to overcome and effectively translate the metallic-polymer nanoconjugates/nanohybrids to the clinic. Additional studies on the biosafety and long-term fate of these nanomaterials are still missing. Moreover, the optimization and scale-up of the production methods are mandatory to decrease batch-to-batch variability. Furthermore, regulatory agencies should create a comprehensive set of guidelines for the translation of metallic nanoparticles to the clinic.
In summary, the metallic-polymer nanoconjugates/nanohybrids have the potential to support the development of more effective and multifunctional all-in-one nanomedicines, with the capacity to diagnose and treat the tumor as well as monitor in real-time its response to the treatment. Such raw potential of metallic-polymer nanoconjugates/nanohybrids presents an area of opportunity for both researchers and industrial partners (e.g., pharmaceutics) to develop a new generation of nanomedicines for cancer therapy. Moreover, it is worthwhile to notice that the physicochemical properties of metallic-polymer nanoconjugates/nanohybrids can also be explored to create innovative solutions in the biomedical field, such as biosensors, antimicrobial agents, and tissue regeneration.

Author Contributions

Conceptualization, A.F.M.; methodology, A.Q.F., N.F., C.F.R. and A.F.M.; software, A.Q.F., C.F.R. and A.F.M.; validation, A.F.M. and I.J.C.; formal analysis, A.F.M. and D.d.M.-D.; investigation, A.Q.F., N.F. and C.F.R.; resources, A.F.M., I.J.C. and D.d.M.-D.; data curation, A.Q.F., N.F. and C.F.R.; writing—original draft preparation, A.Q.F., N.F., C.F.R. and A.F.M.; writing—review and editing, A.F.M., I.J.C. and D.d.M.-D.; visualization, I.J.C. and D.d.M.-D.; supervision, A.F.M. and I.J.C.; project administration, I.J.C.; funding acquisition, I.J.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Foundation for Science and Technology (FCT), through funds from the State Budget, and by the European Regional Development Fund (ERDF), under the Portugal 2020 Program, through the Regional Operational Program of the Center (Centro2020), through the Project with the reference UIDB/00709/2020, CENTRO-01-0145-FEDER-028989 and POCI-01-0145-FEDER-031462. The funder had no role in the decision to publish or in the preparation of the manuscript.

Data Availability Statement

The data presented in this article are available at request from the corresponding author.

Acknowledgments

Carolina F. Rodrigues acknowledges her Ph.D. fellowship from FCT (SFRH/BD/144680/2019). Duarte de Melo-Diogo acknowledges FCT for the financial support given through a Junior Researcher contract (2021.00590.CEECIND). Natanael Fernandes acknowledges the individual fellowship from UBI-Banco Santander/Totta. The funders had no role in the decision to publish or in the preparation of the manuscript.

Conflicts of Interest

The authors declare no financial or commercial conflict of interest.

References

  1. Aghebati-Maleki, A.; Dolati, S.; Ahmadi, M.; Baghbanzhadeh, A.; Asadi, M.; Fotouhi, A.; Yousefi, M.; Aghebati-Maleki, L. Nanoparticles and cancer therapy: Perspectives for application of nanoparticles in the treatment of cancers. J. Cell. Physiol. 2020, 235, 1962–1972. [Google Scholar] [CrossRef] [PubMed]
  2. Goncalves, A.S.C.; Rodrigues, C.F.; Moreira, A.F.; Correia, I.J. Strategies to improve the photothermal capacity of gold-based nanomedicines. Acta Biomater. 2020, 116, 105–137. [Google Scholar] [CrossRef] [PubMed]
  3. Guimaraes, R.S.; Rodrigues, C.F.; Moreira, A.F.; Correia, I.J. Overview of stimuli-responsive mesoporous organosilica nanocarriers for drug delivery. Pharmacol. Res. 2020, 155, 104742. [Google Scholar] [CrossRef] [PubMed]
  4. Huang, N.; Liu, Y.; Fang, Y.; Zheng, S.; Wu, J.; Wang, M.; Zhong, W.; Shi, M.; Xing, M.; Liao, W. Gold Nanoparticles Induce Tumor Vessel Normalization and Impair Metastasis by Inhibiting Endothelial Smad2/3 Signaling. ACS Nano 2020, 14, 7940–7958. [Google Scholar] [CrossRef]
  5. Rodrigues, C.F.; Alves, C.G.; Lima-Sousa, R.; Moreira, A.F.; de Melo-Diogo, D.; Correia, I.J. Inorganic-based drug delivery systems for cancer therapy. In Advances and Avenues in the Development of Novel Carriers for Bioactives and Biological Agents; Academic Press: Cambridge, MA, USA, 2020; pp. 283–316. [Google Scholar]
  6. Liu, X.; Yan, B.; Li, Y.; Ma, X.; Jiao, W.; Shi, K.; Zhang, T.; Chen, S.; He, Y.; Liang, X.; et al. Graphene Oxide-Grafted Magnetic Nanorings Mediated Magnetothermodynamic Therapy Favoring Reactive Oxygen Species-Related Immune Response for Enhanced Antitumor Efficacy. ACS Nano 2020, 14, 1936–1950. [Google Scholar] [CrossRef]
  7. Li, G.; Zhong, X.; Wang, X.; Gong, F.; Lei, H.; Zhou, Y.; Li, C.; Xiao, Z.; Ren, G.; Zhang, L. Titanium carbide nanosheets with defect structure for photothermal-enhanced sonodynamic therapy. Bioact. Mater. 2022, 8, 409–419. [Google Scholar] [CrossRef]
  8. Loh, X.J.; Lee, T.C.; Dou, Q.; Deen, G.R. Utilising inorganic nanocarriers for gene delivery. Biomater. Sci. 2016, 4, 70–86. [Google Scholar] [CrossRef]
  9. Liou, G.Y.; Storz, P. Reactive oxygen species in cancer. Free Radic. Res. 2010, 44, 479–496. [Google Scholar] [CrossRef]
  10. Tricoli, A.; Righettoni, M.; Teleki, A. Semiconductor gas sensors: Dry synthesis and application. Angew. Chem. Int. Ed. 2010, 49, 7632–7659. [Google Scholar] [CrossRef]
  11. Huynh, K.H.; Pham, X.H.; Kim, J.; Lee, S.H.; Chang, H.; Rho, W.Y.; Jun, B.H. Synthesis, Properties, and Biological Applications of Metallic Alloy Nanoparticles. Int. J. Mol. Sci. 2020, 21, 5174. [Google Scholar] [CrossRef]
  12. Amendola, V.; Meneghetti, M. Laser ablation synthesis in solution and size manipulation of noble metal nanoparticles. Phys. Chem. Chem. Phys. 2009, 11, 3805–3821. [Google Scholar] [CrossRef]
  13. Annamalai, J.; Murugan, P.; Ganapathy, D.; Nallaswamy, D.; Atchudan, R.; Arya, S.; Khosla, A.; Barathi, S.; Sundramoorthy, A.K. Synthesis of various dimensional metal organic frameworks (MOFs) and their hybrid composites for emerging applications—A review. Chemosphere 2022, 298, 134184. [Google Scholar] [CrossRef]
  14. Hu, P.; Chen, L.; Kang, X.; Chen, S. Surface Functionalization of Metal Nanoparticles by Conjugated Metal-Ligand Interfacial Bonds: Impacts on Intraparticle Charge Transfer. Acc. Chem. Res. 2016, 49, 2251–2260. [Google Scholar] [CrossRef]
  15. Neha, D.; Momin, M.; Khan, T.; Gharat, S.; Ningthoujam, R.S.; Omri, A. Metallic nanoparticles as drug delivery system for the treatment of cancer. Expert Opin. Drug Deliv. 2021, 18, 1261–1290. [Google Scholar] [CrossRef]
  16. Xu, J.J.; Zhang, W.C.; Guo, Y.W.; Chen, X.Y.; Zhang, Y.N. Metal nanoparticles as a promising technology in targeted cancer treatment. Drug Deliv. 2022, 29, 664–678. [Google Scholar] [CrossRef]
  17. Anselmo, A.C.; Mitragotri, S. A Review of Clinical Translation of Inorganic Nanoparticles. AAPS J. 2015, 17, 1041–1054. [Google Scholar] [CrossRef]
  18. Canaparo, R.; Foglietta, F.; Limongi, T.; Serpe, L. Biomedical Applications of Reactive Oxygen Species Generation by Metal Nanoparticles. Materials 2020, 14, 53. [Google Scholar] [CrossRef]
  19. Gonzalez-Rubio, G.; Guerrero-Martinez, A.; Liz-Marzan, L.M. Reshaping, Fragmentation, and Assembly of Gold Nanoparticles Assisted by Pulse Lasers. Acc. Chem. Res. 2016, 49, 678–686. [Google Scholar] [CrossRef]
  20. Bhatia, S. Natural Polymers vs Synthetic Polymer. In Natural Polymer Drug Delivery Systems; Springer: Cham, Switzerland, 2016; pp. 95–118. [Google Scholar]
  21. Aslan, N.; Ceylan, B.; Koç, M.M.; Findik, F. Metallic nanoparticles as X-ray computed tomography (CT) contrast agents: A review. J. Mol. Struct. 2020, 1219, 128599. [Google Scholar] [CrossRef]
  22. Liu, Y.; Ai, K.; Lu, L. Nanoparticulate X-ray computed tomography contrast agents: From design validation to in vivo applications. Acc. Chem. Res. 2012, 45, 1817–1827. [Google Scholar] [CrossRef]
  23. Cheheltani, R.; Ezzibdeh, R.M.; Chhour, P.; Pulaparthi, K.; Kim, J.; Jurcova, M.; Hsu, J.; Blundell, C.; Litt, H.; Ferrari, V.A.; et al. Tunable, biodegradable gold nanoparticles as contrast agents for computed tomography and photoacoustic imaging. Biomaterials 2016, 102, 87–97. [Google Scholar] [CrossRef] [PubMed]
  24. De La Vega, J.C.; Esquinas, P.L.; Gill, J.K.; Jessa, S.; Gill, B.; Thakur, Y.; Saatchi, K.; Hafeli, U.O. Comparison of Rhenium and Iodine as Contrast Agents in X-ray Imaging. Contrast Media Mol. Imaging 2021, 2021, 1250360. [Google Scholar] [CrossRef] [PubMed]
  25. Berger, M.; Bauser, M.; Frenzel, T.; Hilger, C.S.; Jost, G.; Lauria, S.; Morgenstern, B.; Neis, C.; Pietsch, H.; Sülzle, D.; et al. Hafnium-Based Contrast Agents for X-ray Computed Tomography. Inorg. Chem. 2017, 56, 5757–5761. [Google Scholar] [CrossRef] [PubMed]
  26. Bae, K.T.; McDermott, R.; Gierada, D.S.; Heiken, J.P.; Nolte, M.A.; Takahashi, N.; Hong, C. Gadolinium-enhanced computed tomography angiography in multi-detector row computed tomography. Acad. Radiol. 2004, 11, 61–68. [Google Scholar] [CrossRef]
  27. Werts, M.H.V.; Allix, F.; Francais, O.; Frochot, C.; Griscom, L.; Le Pioufle, B.; Loumaigne, M.; Midelet, J. Manipulation and Optical Detection of Colloidal Functional Plasmonic Nanostructures in Microfluidic Systems. IEEE J. Sel. Top. Quantum Electron. 2014, 20, 102–114. [Google Scholar]
  28. Wang, L.; Hasanzadeh Kafshgari, M.; Meunier, M. Optical Properties and Applications of Plasmonic-Metal Nanoparticles. Adv. Funct. Mater. 2020, 30, 2005400. [Google Scholar] [CrossRef]
  29. Noguez, C. Surface Plasmons on Metal Nanoparticles:  The Influence of Shape and Physical Environment. J. Phys. Chem. C 2007, 111, 3806–3819. [Google Scholar] [CrossRef]
  30. Liz-Marzán, L.M. Tailoring Surface Plasmons through the Morphology and Assembly of Metal Nanoparticles. Langmuir 2006, 22, 32–41. [Google Scholar] [CrossRef]
  31. Liang, J.; Liu, H.; Yu, J.; Zhou, L.; Zhu, J. Plasmon-enhanced solar vapor generation. Nanophotonics 2019, 8, 771–786. [Google Scholar] [CrossRef]
  32. Jeong, Y.; Kook, Y.M.; Lee, K.; Koh, W.G. Metal enhanced fluorescence (MEF) for biosensors: General approaches and a review of recent developments. Biosens. Bioelectron. 2018, 111, 102–116. [Google Scholar] [CrossRef]
  33. Lin, Q.; Jia, M.; Fu, Y.; Li, B.; Dong, Z.; Niu, X.; You, Z. Upper-Critical-Solution-Temperature Polymer Modified Gold Nanorods for Laser Controlled Drug Release and Enhanced Anti-Tumour Therapy. Front. Pharm. 2021, 12, 738630. [Google Scholar] [CrossRef]
  34. de Melo-Diogo, D.; Pais-Silva, C.; Dias, D.R.; Moreira, A.F.; Correia, I.J. Strategies to Improve Cancer Photothermal Therapy Mediated by Nanomaterials. Adv. Healthc. Mater. 2017, 6, 1700073. [Google Scholar] [CrossRef]
  35. Bettaieb, A.; Wrzal, K.P.; Averill-Bates, D.A. Hyperthermia: Cancer Treatment and Beyond. Cancer Treat. Conv. Innov. Approaches 2013, 257–283. [Google Scholar] [CrossRef]
  36. Zhang, Y.; Zhan, X.; Xiong, J.; Peng, S.; Huang, W.; Joshi, R.; Cai, Y.; Liu, Y.; Li, R.; Yuan, K.; et al. Temperature-dependent cell death patterns induced by functionalized gold nanoparticle photothermal therapy in melanoma cells. Sci. Rep. 2018, 8, 8720. [Google Scholar] [CrossRef]
  37. Fernandes, N.; Rodrigues, C.F.; Moreira, A.F.; Correia, I.J. Overview of the application of inorganic nanomaterials in cancer photothermal therapy. Biomater. Sci. 2020, 8, 2990–3020. [Google Scholar] [CrossRef]
  38. Xing, Y.; Cai, Z.; Xu, M.; Ju, W.; Luo, X.; Hu, Y.; Liu, X.; Kang, T.; Wu, P.; Cai, C.; et al. Raman observation of a molecular signaling pathway of apoptotic cells induced by photothermal therapy. Chem. Sci. 2019, 10, 10900–10910. [Google Scholar] [CrossRef]
  39. Jiang, Z.; Li, T.; Cheng, H.; Zhang, F.; Yang, X.; Wang, S.; Zhou, J.; Ding, Y. Nanomedicine potentiates mild photothermal therapy for tumor ablation. Asian J. Pharm. Sci. 2021, 16, 738–761. [Google Scholar] [CrossRef]
  40. Rudakov, G.A.; Tsiberkin, K.B.; Ponomarev, R.S.; Henner, V.K.; Ziolkowska, D.A.; Jasinski, J.B.; Sumanasekera, G. Magnetic properties of transition metal nanoparticles enclosed in carbon nanocages. J. Magn. Magn. Mater. 2019, 472, 34–39. [Google Scholar] [CrossRef]
  41. Soheilian, R.; Choi, Y.S.; David, A.E.; Abdi, H.; Maloney, C.E.; Erb, R.M. Toward Accumulation of Magnetic Nanoparticles into Tissues of Small Porosity. Langmuir 2015, 31, 8267–8274. [Google Scholar] [CrossRef]
  42. Guo, X.; Li, W.; Luo, L.; Wang, Z.; Li, Q.; Kong, F.; Zhang, H.; Yang, J.; Zhu, C.; Du, Y.; et al. External Magnetic Field-Enhanced Chemo-Photothermal Combination Tumor Therapy via Iron Oxide Nanoparticles. ACS Appl. Mater. Interfaces 2017, 9, 16581–16593. [Google Scholar] [CrossRef]
  43. Farzin, A.; Etesami, S.A.; Quint, J.; Memic, A.; Tamayol, A. Magnetic Nanoparticles in Cancer Therapy and Diagnosis. Adv. Healthc. Mater. 2020, 9, e1901058. [Google Scholar] [CrossRef] [PubMed]
  44. Sengul, A.B.; Asmatulu, E. Toxicity of metal and metal oxide nanoparticles: A review. Environ. Chem. Lett. 2020, 18, 1659–1683. [Google Scholar] [CrossRef]
  45. Wu, H.; Yin, J.J.; Wamer, W.G.; Zeng, M.; Lo, Y.M. Reactive oxygen species-related activities of nano-iron metal and nano-iron oxides. J. Food Drug Anal. 2014, 22, 86–94. [Google Scholar] [CrossRef] [PubMed]
  46. Yuan, P.; Ding, X.; Yang, Y.Y.; Xu, Q.-H. Metal Nanoparticles for Diagnosis and Therapy of Bacterial Infection. Adv. Healthc. Mater. 2018, 7, 1701392. [Google Scholar] [CrossRef]
  47. Juarranz, Á.; Jaén, P.; Sanz-Rodríguez, F.; Cuevas, J.; González, S. Photodynamic therapy of cancer. Basic principles and applications. Clin. Transl. Oncol. 2008, 10, 148–154. [Google Scholar] [CrossRef]
  48. Wilson, B.C. Photodynamic Therapy for Cancer: Principles. Can. J. Gastroenterol. 2002, 16, 743109. [Google Scholar] [CrossRef]
  49. Vinković Vrček, I.; Pavičić, I.; Crnković, T.; Jurašin, D.; Babič, M.; Horák, D.; Lovric, M.; Ferhatovic, L.; Curlin, M.; Gajovic, S. Does surface coating of metallic nanoparticles modulate their interference with in vitro assays? RSC Adv. 2015, 5, 70787–70807. [Google Scholar] [CrossRef]
  50. Rajendran, K.; Pujari, L.; Krishnamoorthy, M.; Sen, S.; Dharmaraj, D.; Karuppiah, K.; Ethiraj, K. Toxicological evaluation of biosynthesised hematite nanoparticles in vivo. Colloids Surf. B Biointerfaces 2021, 198, 111475. [Google Scholar] [CrossRef]
  51. García-Álvarez, R.; Hadjidemetriou, M.; Sánchez-Iglesias, A.; Liz-Marzán, L.M.; Kostarelos, K. In vivo formation of protein corona on gold nanoparticles. The effect of their size and shape. Nanoscale 2018, 10, 1256–1264. [Google Scholar] [CrossRef]
  52. Thambiraj, S.; Vijayalakshmi, R.; Ravi Shankaran, D. An effective strategy for development of docetaxel encapsulated gold nanoformulations for treatment of prostate cancer. Sci. Rep. 2021, 11, 2808. [Google Scholar] [CrossRef]
  53. Hada, A.M.; Craciun, A.M.; Focsan, M.; Borlan, R.; Soritau, O.; Todea, M.; Astilean, S. Folic acid functionalized gold nanoclusters for enabling targeted fluorescence imaging of human ovarian cancer cells. Talanta 2021, 225, 121960. [Google Scholar] [CrossRef]
  54. Kim, H.S.; Lee, S.J.; Lee, D.Y. Milk protein-shelled gold nanoparticles with gastrointestinally active absorption for aurotherapy to brain tumor. Bioact. Mater. 2022, 8, 35–48. [Google Scholar] [CrossRef]
  55. Mapanao, A.K.; Santi, M.; Voliani, V. Combined chemo-photothermal treatment of three-dimensional head and neck squamous cell carcinomas by gold nano-architectures. J. Colloid Interface Sci. 2021, 582, 1003–1011. [Google Scholar] [CrossRef]
  56. You, Y.H.; Lin, Y.F.; Nirosha, B.; Chang, H.T.; Huang, Y.F. Polydopamine-coated gold nanostar for combined antitumor and antiangiogenic therapy in multidrug-resistant breast cancer. Nanotheranostics 2019, 3, 266–283. [Google Scholar] [CrossRef]
  57. Sheng, Y.; Liu, C.; Yuan, Y.; Tao, X.; Yang, F.; Shan, X.; Zhou, H.; Xu, F. Long-circulating polymeric nanoparticles bearing a combinatorial coating of PEG and water-soluble chitosan. Biomaterials 2009, 30, 2340–2348. [Google Scholar] [CrossRef]
  58. Li, B.; Xie, J.; Yuan, Z.; Jain, P.; Lin, X.; Wu, K.; Jiang, S. Mitigation of Inflammatory Immune Responses with Hydrophilic Nanoparticles. Angew. Chem. Int. Ed. 2018, 57, 4527–4531. [Google Scholar] [CrossRef]
  59. Lowe, S.; O’Brien-Simpson, N.M.; Connal, L.A. Antibiofouling polymer interfaces: Poly(ethylene glycol) and other promising candidates. Polym. Chem. 2015, 6, 198–212. [Google Scholar] [CrossRef]
  60. Yu, Y.; Luan, Y.; Dai, W. Dynamic process, mechanisms, influencing factors and study methods of protein corona formation. Int. J. Biol. Macromol. 2022, 205, 731–739. [Google Scholar] [CrossRef]
  61. Feng, W.; Zhu, S.; Ishihara, K.; Brash, J.L. Protein resistant surfaces: Comparison of acrylate graft polymers bearing oligo-ethylene oxide and phosphorylcholine side chains. Biointerphases 2006, 1, 50. [Google Scholar] [CrossRef]
  62. He, M.; Gao, K.; Zhou, L.; Jiao, Z.; Wu, M.; Cao, J.; You, X.; Cai, Z.; Su, Y.; Jiang, Z. Zwitterionic materials for antifouling membrane surface construction. Acta Biomater. 2016, 40, 142–152. [Google Scholar] [CrossRef]
  63. Zhang, Y.; Liu, Y.; Ren, B.; Zhang, D.; Xie, S.; Chang, Y.; Yang, J.; Wu, J.; Xu, L.; Zheng, J. Fundamentals and applications of zwitterionic antifouling polymers. J. Phys. D Appl. Phys. 2019, 52, 403001. [Google Scholar] [CrossRef]
  64. Liu, X.; Li, H.; Chen, Y.; Jin, Q.; Ren, K.; Ji, J. Mixed-Charge Nanoparticles for Long Circulation, Low Reticuloendothelial System Clearance, and High Tumor Accumulation. Adv. Healthc. Mater. 2014, 3, 1439–1447. [Google Scholar] [CrossRef]
  65. Wu, L.; Lin, B.; Yang, H.; Chen, J.; Mao, Z.; Wang, W.; Gao, C. Enzyme-responsive multifunctional peptide coating of gold nanorods improves tumor targeting and photothermal therapy efficacy. Acta Biomater. 2019, 86, 363–372. [Google Scholar] [CrossRef] [PubMed]
  66. Li, W.; Cao, Z.; Yu, L.; Huang, Q.; Zhu, D.; Lu, C.; Lu, A.; Liu, Y. Hierarchical drug release designed Au @PDA-PEG-MTX NPs for targeted delivery to breast cancer with combined photothermal-chemotherapy. J. Nanobiotechnol. 2021, 19, 143. [Google Scholar] [CrossRef] [PubMed]
  67. Sathiyaseelan, A.; Saravanakumar, K.; Mariadoss, A.V.A.; Wang, M.H. pH-controlled nucleolin targeted release of dual drug from chitosan-gold based aptamer functionalized nano drug delivery system for improved glioblastoma treatment. Carbohydr. Polym. 2021, 262, 117907. [Google Scholar] [CrossRef] [PubMed]
  68. Feng, Y.; Cheng, Y.; Chang, Y.; Jian, H.; Zheng, R.; Wu, X.; Xu, K.; Wang, L.; Ma, X.; Li, X.; et al. Time-staggered delivery of erlotinib and doxorubicin by gold nanocages with two smart polymers for reprogrammable release and synergistic with photothermal therapy. Biomaterials 2019, 217, 119327. [Google Scholar] [CrossRef]
  69. Wang, N.; Shi, J.; Wu, C.; Chu, W.; Tao, W.; Li, W.; Yuan, X. Design of DOX-GNRs-PNIPAM@PEG-PLA Micelle With Temperature and Light Dual-Function for Potent Melanoma Therapy. Front. Chem. 2020, 8, 599740. [Google Scholar] [CrossRef]
  70. Li, D.; Zhang, R.; Liu, G.; Kang, Y.; Wu, J. Redox-Responsive Self-Assembled Nanoparticles for Cancer Therapy. Adv. Healthc. Mater. 2020, 9, 2000605. [Google Scholar] [CrossRef]
  71. Shi, Y.; van der Meel, R.; Chen, X.; Lammers, T. The EPR effect and beyond: Strategies to improve tumor targeting and cancer nanomedicine treatment efficacy. Theranostics 2020, 10, 7921–7924. [Google Scholar] [CrossRef]
  72. Acharya, S.; Sahoo, S.K. PLGA nanoparticles containing various anticancer agents and tumour delivery by EPR effect. Adv. Drug Deliv. Rev. 2011, 63, 170–183. [Google Scholar] [CrossRef]
  73. Yucel, O.; Sengelen, A.; Emik, S.; Onay-Ucar, E.; Arda, N.; Gurdag, G. Folic acid-modified methotrexate-conjugated gold nanoparticles as nano-sized trojans for drug delivery to folate receptor-positive cancer cells. Nanotechnology 2020, 31, 355101. [Google Scholar] [CrossRef]
  74. Mulens-Arias, V.; Nicolas-Boluda, A.; Pinto, A.; Balfourier, A.; Carn, F.; Silva, A.K.A.; Pocard, M.; Gazeau, F. Tumor-Selective Immune-Active Mild Hyperthermia Associated with Chemotherapy in Colon Peritoneal Metastasis by Photoactivation of Fluorouracil-Gold Nanoparticle Complexes. ACS Nano 2021, 15, 3330–3348. [Google Scholar] [CrossRef]
  75. Guo, J.; Rahme, K.; He, Y.; Li, L.L.; Holmes, J.D.; O’Driscoll, C.M. Gold nanoparticles enlighten the future of cancer theranostics. Int. J. Nanomed. 2017, 12, 6131–6152. [Google Scholar] [CrossRef]
  76. Cole, L.E.; Ross, R.D.; Tilley, J.M.R.; Vargo-Gogola, T.; Roeder, R.K. Gold nanoparticles as contrast agents in x-ray imaging and computed tomography. Nanomedicine 2015, 10, 321–341. [Google Scholar] [CrossRef]
  77. Xi, D.; Dong, S.; Meng, X.; Lu, Q.; Meng, L.; Ye, J. Gold nanoparticles as computerized tomography (CT) contrast agents. RSC Adv. 2012, 2, 12515. [Google Scholar] [CrossRef]
  78. Capek, I. Polymer decorated gold nanoparticles in nanomedicine conjugates. Adv. Colloid Interface Sci. 2017, 249, 386–399. [Google Scholar] [CrossRef]
  79. Guo, C.; Hall, G.N.; Addison, J.B.; Yarger, J.L. Gold nanoparticle-doped silk film as biocompatible SERS substrate. RSC Adv. 2015, 5, 1937–1942. [Google Scholar] [CrossRef]
  80. Fernandes, N.; Rodrigues, C.F.; de Melo-Diogo, D.; Correia, I.J.; Moreira, A.F. Optimization of the GSH-Mediated Formation of Mesoporous Silica-Coated Gold Nanoclusters for NIR Light-Triggered Photothermal Applications. Nanomaterials 2021, 11, 1946. [Google Scholar] [CrossRef]
  81. Fernandes, J.; Kang, S. Numerical Study on the Surface Plasmon Resonance Tunability of Spherical and Non-Spherical Core-Shell Dimer Nanostructures. Nanomaterials 2021, 11, 1728. [Google Scholar] [CrossRef]
  82. Xu, H.; Käll, M. Modeling the optical response of nanoparticle-based surface plasmon resonance sensors. Sens. Actuators B Chem. 2002, 87, 244–249. [Google Scholar] [CrossRef]
  83. Bouhelier, A.; Bachelot, R.; Lerondel, G.; Kostcheev, S.; Royer, P.; Wiederrecht, G.P. Surface plasmon characteristics of tunable photoluminescence in single gold nanorods. Phys. Rev. Lett 2005, 95, 267405. [Google Scholar] [CrossRef] [PubMed]
  84. Chandrasekaran, R.; Lee, A.S.W.; Yap, L.W.; Jans, D.A.; Wagstaff, K.M.; Cheng, W. Tumor cell-specific photothermal killing by SELEX-derived DNA aptamer-targeted gold nanorods. Nanoscale 2016, 8, 187–196. [Google Scholar] [CrossRef] [PubMed]
  85. Shi, W.; Casas, J.; Venkataramasubramani, M.; Tang, L. Synthesis and Characterization of Gold Nanoparticles with Plasmon Absorbance Wavelength Tunable from Visible to Near Infrared Region. ISRN Nanomater. 2012, 2012, 659043. [Google Scholar] [CrossRef]
  86. Jacinto, T.A.; Rodrigues, C.F.; Moreira, A.F.; Miguel, S.P.; Costa, E.C.; Ferreira, P.; Correia, I.J. Hyaluronic acid and vitamin E polyethylene glycol succinate functionalized gold-core silica shell nanorods for cancer targeted photothermal therapy. Colloids Surf. B Biointerfaces 2020, 188, 110778. [Google Scholar] [CrossRef]
  87. F Rodrigues, C.; Fernandes, N.; de Melo-Diogo, D.; Ferreira, P.; J Correia, I.; F Moreira, A. HA/PEI-coated acridine orange-loaded gold-core silica shell nanorods for cancer-targeted photothermal and chemotherapy. Nanomedicine 2021, 16, 2569–2586. [Google Scholar] [CrossRef]
  88. Pan, Y.; Ma, X.; Liu, C.; Xing, J.; Zhou, S.; Parshad, B.; Schwerdtle, T.; Li, W.; Wu, A.; Haag, R. Retinoic Acid-Loaded Dendritic Polyglycerol-Conjugated Gold Nanostars for Targeted Photothermal Therapy in Breast Cancer Stem Cells. ACS Nano 2021, 15, 15069–15084. [Google Scholar] [CrossRef]
  89. Tan, H.; Hou, N.; Liu, Y.; Liu, B.; Cao, W.; Zheng, D.; Li, W.; Liu, Y.; Xu, B.; Wang, Z.; et al. CD133 antibody targeted delivery of gold nanostars loading IR820 and docetaxel for multimodal imaging and near-infrared photodynamic/photothermal/chemotherapy against castration resistant prostate cancer. Nanomedicine 2020, 27, 102192. [Google Scholar] [CrossRef]
  90. Cheng, Y.; Bao, D.; Chen, X.; Wu, Y.; Wei, Y.; Wu, Z.; Li, F.; Piao, J.G. Microwave-triggered/HSP-targeted gold nano-system for triple-negative breast cancer photothermal therapy. Int. J. Pharm. 2021, 593, 120162. [Google Scholar] [CrossRef]
  91. Peng, C.; Zheng, L.; Chen, Q.; Shen, M.; Guo, R.; Wang, H.; Cao, X.; Zhang, G.; Shi, X. PEGylated dendrimer-entrapped gold nanoparticles for in vivo blood pool and tumor imaging by computed tomography. Biomaterials 2012, 33, 1107–1119. [Google Scholar] [CrossRef]
  92. Gu, W.; Zhang, Q.; Zhang, T.; Li, Y.; Xiang, J.; Peng, R.; Liu, J. Hybrid polymeric nano-capsules loaded with gold nanoclusters and indocyanine green for dual-modal imaging and photothermal therapy. J. Mater. Chem. B 2016, 4, 910–919. [Google Scholar] [CrossRef]
  93. Alkhayal, A.; Fathima, A.; Alhasan, A.H.; Alsharaeh, E.H. PEG Coated Fe3O4/RGO Nano-Cube-Like Structures for Cancer Therapy via Magnetic Hyperthermia. Nanomaterials 2021, 11, 2398. [Google Scholar] [CrossRef]
  94. Ebrahiminezhad, A.; Zare-Hoseinabadi, A.; Sarmah, A.K.; Taghizadeh, S.; Ghasemi, Y.; Berenjian, A. Plant-Mediated Synthesis and Applications of Iron Nanoparticles. Mol. Biotechnol. 2018, 60, 154–168. [Google Scholar] [CrossRef]
  95. Zhi, D.; Yang, T.; Yang, J.; Fu, S.; Zhang, S. Targeting strategies for superparamagnetic iron oxide nanoparticles in cancer therapy. Acta Biomater. 2020, 102, 13–34. [Google Scholar] [CrossRef]
  96. Palanisamy, S.; Wang, Y.M. Superparamagnetic iron oxide nanoparticulate system: Synthesis, targeting, drug delivery and therapy in cancer. Dalton Trans. 2019, 48, 9490–9515. [Google Scholar] [CrossRef]
  97. Habra, K.; McArdle, S.E.B.; Morris, R.H.; Cave, G.W.V. Synthesis and Functionalisation of Superparamagnetic Nano-Rods towards the Treatment of Glioblastoma Brain Tumours. Nanomaterials 2021, 11, 2157. [Google Scholar] [CrossRef]
  98. Vangijzegem, T.; Stanicki, D.; Laurent, S. Magnetic iron oxide nanoparticles for drug delivery: Applications and characteristics. Expert Opin. Drug Deliv. 2019, 16, 69–78. [Google Scholar] [CrossRef]
  99. Liu, J.F.; Jang, B.; Issadore, D.; Tsourkas, A. Use of magnetic fields and nanoparticles to trigger drug release and improve tumor targeting. WIREs Nanomed. Nanobiotechnol. 2019, 11, e1571. [Google Scholar] [CrossRef]
  100. Shah, R.R.; Davis, T.P.; Glover, A.L.; Nikles, D.E.; Brazel, C.S. Impact of magnetic field parameters and iron oxide nanoparticle properties on heat generation for use in magnetic hyperthermia. J. Magn. Magn. Mater. 2015, 387, 96–106. [Google Scholar] [CrossRef]
  101. Obaidat, I.M.; Issa, B.; Haik, Y. Magnetic Properties of Magnetic Nanoparticles for Efficient Hyperthermia. Nanomaterials 2015, 5, 63–89. [Google Scholar] [CrossRef]
  102. Boyer, C.; Whittaker, M.R.; Bulmus, V.; Liu, J.; Davis, T.P. The design and utility of polymer-stabilized iron-oxide nanoparticles for nanomedicine applications. NPG Asia Mater. 2010, 2, 23–30. [Google Scholar] [CrossRef]
  103. Xu, X.; Zhou, X.; Xiao, B.; Xu, H.; Hu, D.; Qian, Y.; Hu, H.; Zhou, Z.; Liu, X.; Gao, J.; et al. Glutathione-Responsive Magnetic Nanoparticles for Highly Sensitive Diagnosis of Liver Metastases. Nano Lett. 2021, 21, 2199–2206. [Google Scholar] [CrossRef]
  104. Xiao, Z.; You, Y.; Liu, Y.; He, L.; Zhang, D.; Cheng, Q.; Wang, D.; Chen, T.; Shi, C.; Luo, L. NIR-Triggered Blasting Nanovesicles for Targeted Multimodal Image-Guided Synergistic Cancer Photothermal and Chemotherapy. ACS Appl. Mater. Interfaces 2021, 13, 35376–35388. [Google Scholar] [CrossRef] [PubMed]
  105. Chen, L.; Wu, Y.; Wu, H.; Li, J.; Xie, J.; Zang, F.; Ma, M.; Gu, N.; Zhang, Y. Magnetic targeting combined with active targeting of dual-ligand iron oxide nanoprobes to promote the penetration depth in tumors for effective magnetic resonance imaging and hyperthermia. Acta Biomater. 2019, 96, 491–504. [Google Scholar] [CrossRef] [PubMed]
  106. Zhang, Y.; Hu, H.; Tang, W.; Zhang, Q.; Li, M.; Jin, H.; Huang, Z.; Cui, Z.; Xu, J.; Wang, K.; et al. A multifunctional magnetic nanosystem based on “two strikes” effect for synergistic anticancer therapy in triple-negative breast cancer. J. Control. Release 2020, 322, 401–415. [Google Scholar] [CrossRef] [PubMed]
  107. Zheng, D.; Wan, C.; Yang, H.; Xu, L.; Dong, Q.; Du, C.; Du, J.; Li, F. Her2-Targeted Multifunctional Nano-Theranostic Platform Mediates Tumor Microenvironment Remodeling and Immune Activation for Breast Cancer Treatment. Int. J. Nanomed. 2020, 15, 10007–10028. [Google Scholar] [CrossRef]
  108. He, Y.; Wang, M.; Fu, M.; Yuan, X.; Luo, Y.; Qiao, B.; Cao, J.; Wang, Z.; Hao, L.; Yuan, G. Iron(II) phthalocyanine Loaded and AS1411 Aptamer Targeting Nanoparticles: A Nanocomplex for Dual Modal Imaging and Photothermal Therapy of Breast Cancer. Int. J. Nanomed. 2020, 15, 5927–5949. [Google Scholar] [CrossRef]
  109. Ding, X.; Jiang, W.; Dong, L.; Hong, C.; Luo, Z.; Hu, Y.; Cai, K. Redox-responsive magnetic nanovectors self-assembled from amphiphilic polymer and iron oxide nanoparticles for a remotely targeted delivery of paclitaxel. J. Mater. Chem. B 2021, 9, 6037–6043. [Google Scholar] [CrossRef]
  110. Lin, C.H.; Chen, Y.C.; Huang, P.I. Preparation of Multifunctional Dopamine-Coated Zerovalent Iron/Reduced Graphene Oxide for Targeted Phototheragnosis in Breast Cancer. Nanomaterials 2020, 10, 1957. [Google Scholar] [CrossRef]
  111. Yun, B.; Zhu, H.; Yuan, J.; Sun, Q.; Li, Z. Synthesis, modification and bioapplications of nanoscale copper chalcogenides. J. Mater. Chem. B 2020, 8, 4778–4812. [Google Scholar] [CrossRef]
  112. Zhou, M.; Tian, M.; Li, C. Copper-Based Nanomaterials for Cancer Imaging and Therapy. Bioconjugate Chem. 2016, 27, 1188–1199. [Google Scholar] [CrossRef]
  113. Rubilar, O.; Rai, M.; Tortella, G.; Diez, M.C.; Seabra, A.B.; Duran, N. Biogenic nanoparticles: Copper, copper oxides, copper sulphides, complex copper nanostructures and their applications. Biotechnol. Lett. 2013, 35, 1365–1375. [Google Scholar] [CrossRef]
  114. Tian, H.; Zhang, M.; Jin, G.; Jiang, Y.; Luan, Y. Cu-MOF chemodynamic nanoplatform via modulating glutathione and H2O2 in tumor microenvironment for amplified cancer therapy. J. Colloid Interface Sci. 2021, 587, 358–366. [Google Scholar] [CrossRef]
  115. Wang, S.; Riedinger, A.; Li, H.; Fu, C.; Liu, H.; Li, L.; Liu, T.; Tan, L.; Barthel, M.J.; Pugliese, G.; et al. Plasmonic Copper Sulfide Nanocrystals Exhibiting Near-Infrared Photothermal and Photodynamic Therapeutic Effects. ACS Nano 2015, 9, 1788–1800. [Google Scholar] [CrossRef]
  116. Egorova, K.S.; Ananikov, V.P. Which Metals are Green for Catalysis? Comparison of the Toxicities of Ni, Cu, Fe, Pd, Pt, Rh, and Au Salts. Angew. Chem. Int. Ed. 2016, 55, 12150–12162. [Google Scholar] [CrossRef]
  117. Letelier, M.E.; Sánchez-Jofré, S.; Peredo-Silva, L.; Cortés-Troncoso, J.; Aracena-Parks, P. Mechanisms underlying iron and copper ions toxicity in biological systems: Pro-oxidant activity and protein-binding effects. Chem. Biol. Interact. 2010, 188, 220–227. [Google Scholar] [CrossRef]
  118. Li, W.; Zamani, R.; Gil, P.R.; Pelaz, B.; Ibáñez, M.; Cadavid, D.; Shavel, A.; Alvarez-Puebla, R.A.; Parak, W.J.; Arbiol, J.; et al. CuTe Nanocrystals: Shape and Size Control, Plasmonic Properties, and Use as SERS Probes and Photothermal Agents. J. Am. Chem. Soc. 2013, 135, 7098–7101. [Google Scholar] [CrossRef]
  119. Li, L.; Rashidi, L.H.; Yao, M.; Ma, L.; Chen, L.; Zhang, J.; Zhang, Y.; Chen, W. CuS nanoagents for photodynamic and photothermal therapies: Phenomena and possible mechanisms. Photodiagn. Photodyn. 2017, 19, 5–14. [Google Scholar] [CrossRef]
  120. Shi, H.; Yan, R.; Wu, L.; Sun, Y.; Liu, S.; Zhou, Z.; He, J.; Ye, D. Tumor-targeting CuS nanoparticles for multimodal imaging and guided photothermal therapy of lymph node metastasis. Acta Biomater. 2018, 72, 256–265. [Google Scholar] [CrossRef]
  121. Xu, J.; Shi, R.; Chen, G.; Dong, S.; Yang, P.; Zhang, Z.; Niu, N.; Gai, S.; He, F.; Fu, Y.; et al. All-in-One Theranostic Nanomedicine with Ultrabright Second Near-Infrared Emission for Tumor-Modulated Bioimaging and Chemodynamic/Photodynamic Therapy. ACS Nano 2020, 14, 9613–9625. [Google Scholar] [CrossRef]
  122. Liang, S.; Deng, X.; Chang, Y.; Sun, C.; Shao, S.; Xie, Z.; Xiao, X.; Ma, P.; Zhang, H.; Cheng, Z.; et al. Intelligent Hollow Pt-CuS Janus Architecture for Synergistic Catalysis-Enhanced Sonodynamic and Photothermal Cancer Therapy. Nano Lett. 2019, 19, 4134–4145. [Google Scholar] [CrossRef]
  123. Poudel, K.; Thapa, R.K.; Gautam, M.; Ou, W.; Soe, Z.C.; Gupta, B.; Ruttala, H.B.; Thuy, H.N.; Dai, P.C.; Jeong, J.-H.; et al. Multifaceted NIR-responsive polymer-peptide-enveloped drug-loaded copper sulfide nanoplatform for chemo-phototherapy against highly tumorigenic prostate cancer. Nanomedicine 2019, 21, 102042. [Google Scholar] [CrossRef] [PubMed]
  124. Maor, I.; Asadi, S.; Korganbayev, S.; Dahis, D.; Shamay, Y.; Schena, E.; Azhari, H.; Saccomandi, P.; Weitz, I.S. Laser-induced thermal response and controlled release of copper oxide nanoparticles from multifunctional polymeric nanocarriers. Sci. Technol. Adv. Mater. 2021, 22, 218–233. [Google Scholar] [CrossRef] [PubMed]
  125. Xu, R.; Zhang, K.; Liang, J.; Gao, F.; Li, J.; Guan, F. Hyaluronic acid/polyethyleneimine nanoparticles loaded with copper ion and disulfiram for esophageal cancer. Carbohydr. Polym. 2021, 261, 117846. [Google Scholar] [CrossRef] [PubMed]
  126. Wu, Z.; Zhang, P.; Wang, P.; Wang, Z.; Luo, X. Using copper sulfide nanoparticles as cross-linkers of tumor microenvironment responsive polymer micelles for cancer synergistic photo-chemotherapy. Nanoscale 2021, 13, 3723–3736. [Google Scholar] [CrossRef]
  127. Xiao, Z.; Zuo, W.; Chen, L.; Wu, L.; Liu, N.; Liu, J.; Jin, Q.; Zhao, Y.; Zhu, X. H2O2 Self-Supplying and GSH-Depleting Nanoplatform for Chemodynamic Therapy Synergetic Photothermal/Chemotherapy. ACS Appl. Mater. Interfaces 2021, 13, 43925–43936. [Google Scholar] [CrossRef]
  128. Cai, X.; Xie, Z.; Ding, B.; Shao, S.; Liang, S.; Pang, M.; Lin, J. Monodispersed Copper(I)-Based Nano Metal-Organic Framework as a Biodegradable Drug Carrier with Enhanced Photodynamic Therapy Efficacy. Adv. Sci. 2019, 6, 1900848. [Google Scholar] [CrossRef]
  129. Fang, X.L.; Akrofi, R.; Yang, H.; Chen, Q.Y. The NIR inspired nano-CuSMn(II) composites for lactate and glycolysis attenuation. Colloids Surf. B Biointerfaces 2019, 181, 728–733. [Google Scholar] [CrossRef]
  130. Yu, X.; Yu, J.; Cheng, B.; Huang, B. One-Pot Template-Free Synthesis of Monodisperse Zinc Sulfide Hollow Spheres and Their Photocatalytic Properties. Chem. A Eur. J. 2009, 15, 6731–6739. [Google Scholar] [CrossRef]
  131. Wang, C.-C.; Wang, S.; Xia, Q.; He, W.; Yin, J.-J.; Fu, P.P.; Li, J.-H. Phototoxicity of Zinc Oxide Nanoparticles in HaCaT Keratinocytes-Generation of Oxidative DNA Damage During UVA and Visible Light Irradiation. J. Nanosci. Nanotechnol. 2013, 13, 3880–3888. [Google Scholar] [CrossRef]
  132. Akhtar, M.J.; Ahamed, M.; Kumar, S.; Khan, M.M.; Ahmad, J.; Alrokayan, S.A. Zinc oxide nanoparticles selectively induce apoptosis in human cancer cells through reactive oxygen species. Int. J. Nanomed. 2012, 7, 845–857. [Google Scholar]
  133. Song, T.; Qu, Y.; Ren, Z.; Yu, S.; Sun, M.; Yu, X.; Yu, X. Synthesis and Characterization of Polyvinylpyrrolidone-Modified ZnO Quantum Dots and Their In Vitro Photodynamic Tumor Suppressive Action. Int. J. Mol. Sci. 2021, 22, 8106. [Google Scholar] [CrossRef]
  134. Riddell, I.A.; Lippard, S.J. Cisplatin and Oxaliplatin: Our Current Understanding of Their Actions. Met. Ions Life Sci. 2018, 18, 1–42. [Google Scholar]
  135. Asharani, P.V.; Xinyi, N.; Hande, M.P.; Valiyaveettil, S. DNA damage and p53-mediated growth arrest in human cells treated with platinum nanoparticles. Nanomedicine 2009, 5, 51–64. [Google Scholar] [CrossRef]
  136. Cao, H.; Yang, Y.; Liang, M.; Ma, Y.; Sun, N.; Gao, X.; Li, J. Pt@polydopamine nanoparticles as nanozymes for enhanced photodynamic and photothermal therapy. Chem. Commun. 2021, 57, 255–258. [Google Scholar] [CrossRef]
  137. Pedone, D.; Moglianetti, M.; De Luca, E.; Bardi, G.; Pompa, P.P. Platinum nanoparticles in nanobiomedicine. Chem. Soc. Rev. 2017, 46, 4951–4975. [Google Scholar] [CrossRef]
  138. Chen, T.; Gu, T.; Cheng, L.; Li, X.; Han, G.; Liu, Z. Porous Pt nanoparticles loaded with doxorubicin to enable synergistic Chemo-/Electrodynamic Therapy. Biomaterials 2020, 255, 120202. [Google Scholar] [CrossRef]
  139. Zhu, Y.; Li, W.; Zhao, X.; Zhou, Z.; Wang, Y.; Cheng, Y.; Huang, Q.; Zhang, Q. Hyaluronic Acid-Encapsulated Platinum Nanoparticles for Targeted Photothermal Therapy of Breast Cancer. J. Biomed. Nanotechnol. 2017, 13, 1457–1467. [Google Scholar] [CrossRef]
  140. Zhang, X.F.; Liu, Z.G.; Shen, W.; Gurunathan, S. Silver Nanoparticles: Synthesis, Characterization, Properties, Applications, and Therapeutic Approaches. Int. J. Mol. Sci. 2016, 17, 1534. [Google Scholar] [CrossRef]
  141. Kim, S.; Ryu, D.-Y. Silver nanoparticle-induced oxidative stress, genotoxicity and apoptosis in cultured cells and animal tissues. J. Appl. Toxicol. 2013, 33, 78–89. [Google Scholar] [CrossRef]
  142. Holmila, R.J.; Vance, S.A.; King, S.B.; Tsang, A.W.; Singh, R.; Furdui, C.M. Silver Nanoparticles Induce Mitochondrial Protein Oxidation in Lung Cells Impacting Cell Cycle and Proliferation. Antioxidants 2019, 8, 552. [Google Scholar] [CrossRef]
  143. Park, T.; Lee, S.; Amatya, R.; Cheong, H.; Moon, C.; Kwak, H.D.; Min, K.A.; Shin, M.C. ICG-Loaded PEGylated BSA-Silver Nanoparticles for Effective Photothermal Cancer Therapy. Int. J. Nanomed. 2020, 15, 5459–5471. [Google Scholar] [CrossRef]
  144. Zhang, J.; Zhao, B.; Chen, S.; Wang, Y.; Xiao, H. Near-Infrared Light Irradiation Induced Mild Hyperthermia Enhances Glutathione Depletion and DNA Interstrand Cross-Link Formation for Efficient Chemotherapy. ACS Nano 2020, 14, 14831–14845. [Google Scholar] [CrossRef]
  145. Awasthi, P.; An, X.; Xiang, J.; Kalva, N.; Shen, Y.; Li, C. Facile synthesis of noncytotoxic PEGylated dendrimer encapsulated silver sulfide quantum dots for NIR-II biological imaging. Nanoscale 2020, 12, 5678–5684. [Google Scholar] [CrossRef]
  146. Chong, Y.; Huang, J.; Xu, X.; Yu, C.; Ning, X.; Fan, S.; Zhang, Z. Hyaluronic Acid-Modified Au–Ag Alloy Nanoparticles for Radiation/Nanozyme/Ag+ Multimodal Synergistically Enhanced Cancer Therapy. Bioconjugate Chem. 2020, 31, 1756–1765. [Google Scholar] [CrossRef]
  147. Zhang, X.-S.; Xuan, Y.; Yang, X.-Q.; Cheng, K.; Zhang, R.-Y.; Li, C.; Tan, F.; Cao, Y.-C.; Song, X.-L. A multifunctional targeting probe with dual-mode imaging and photothermal therapy used in vivo. J. Nanobiotechnology 2018, 16, 42. [Google Scholar] [CrossRef]
  148. Liu, M.; Peng, Y.; Nie, Y.; Liu, P.; Hu, S.; Ding, J.; Zhou, W. Co-delivery of doxorubicin and DNAzyme using ZnO@polydopamine core-shell nanocomposites for chemo/gene/photothermal therapy. Acta Biomater. 2020, 110, 242–253. [Google Scholar] [CrossRef]
  149. Sun, Y.; Yan, C.; Xie, J.; Yan, D.; Hu, K.; Huang, S.; Liu, J.; Zhang, Y.; Gu, N.; Xiong, F. High-Performance Worm-like Mn-Zn Ferrite Theranostic Nanoagents and the Application on Tumor Theranostics. ACS Appl. Mater. Interfaces 2019, 11, 29536–29548. [Google Scholar] [CrossRef] [PubMed]
  150. Thakur, N.S.; Patel, G.; Kushwah, V.; Jain, S.; Banerjee, U.C.; Thakur, N.S. Facile development of biodegradable polymer-based nanotheranostics: Hydrophobic photosensitizers delivery, fluorescence imaging and photodynamic therapy. J. Photochem. Photobiol. B 2019, 193, 39–50. [Google Scholar] [CrossRef] [PubMed]
  151. Anselmo, A.C.; Mitragotri, S. Nanoparticles in the clinic: An update post COVID-19 vaccines. Bioeng. Transl. Med. 2021, 6, e10246. [Google Scholar] [CrossRef] [PubMed]
  152. Rodallec, A.; Benzekry, S.; Lacarelle, B.; Ciccolini, J.; Fanciullino, R. Pharmacokinetics variability: Why nanoparticles are not just magic-bullets in oncology. Crit. Rev. Oncol. Hematol. 2018, 129, 1–12. [Google Scholar] [CrossRef] [PubMed]
  153. Wilhelm, S.; Tavares, A.J.; Dai, Q.; Ohta, S.; Audet, J.; Dvorak, H.F.; Chan, W.C.W. Analysis of nanoparticle delivery to tumours. Nat. Rev. Mater. 2016, 1, 16014. [Google Scholar] [CrossRef]
  154. Shi, J.; Kantoff, P.W.; Wooster, R.; Farokhzad, O.C. Cancer nanomedicine: Progress, challenges and opportunities. Nat. Rev. Cancer 2017, 17, 20–37. [Google Scholar] [CrossRef]
  155. Rastinehad, A.R.; Anastos, H.; Wajswol, E.; Winoker, J.S.; Sfakianos, J.P.; Doppalapudi, S.K.; Carrick, M.R.; Knauer, C.J.; Taouli, B.; Lewis, S.C. Gold nanoshell-localized photothermal ablation of prostate tumors in a clinical pilot device study. Proc. Natl. Acad. Sci. USA 2019, 116, 18590–18596. [Google Scholar] [CrossRef] [Green Version]
  156. Bayda, S.; Hadla, M.; Palazzolo, S.; Riello, P.; Corona, G.; Toffoli, G.; Rizzolio, F. Inorganic Nanoparticles for Cancer Therapy: A Transition from Lab to Clinic. Curr. Med. Chem. 2018, 25, 4269–4303. [Google Scholar] [CrossRef]
  157. Kumthekar, P.; Ko, C.H.; Paunesku, T.; Dixit, K.; Sonabend, A.M.; Bloch, O.; Tate, M.; Schwartz, M.; Zuckerman, L.; Lezon, R. A first-in-human phase 0 clinical study of RNA interference-based spherical nucleic acids in patients with recurrent glioblastoma. Sci. Transl. Med. 2021, 13, eabb3945. [Google Scholar] [CrossRef]
  158. Kumthekar, P.; Rademaker, A.; Ko, C.; Dixit, K.; Schwartz, M.A.; Sonabend, A.M.; Sharp, L.; Lukas, R.V.; Stupp, R.; Horbinski, C.; et al. A phase 0 first-in-human study using NU-0129: A gold base spherical nucleic acid (SNA) nanoconjugate targeting BCL2L12 in recurrent glioblastoma patients. J. Clin. Oncol. 2019, 37 (Suppl. S15), 3012. [Google Scholar] [CrossRef]
Figure 1. Overview of the properties of metallic nanoparticles, advantages of the polymers’ inclusion, and representation of the most common structural organizations of the metal-polymer nanohybrids/complexes.
Figure 1. Overview of the properties of metallic nanoparticles, advantages of the polymers’ inclusion, and representation of the most common structural organizations of the metal-polymer nanohybrids/complexes.
Nanomaterials 12 03166 g001
Figure 2. In vivo evaluation of the pAAu-Erl plus pNAu-Dox antitumoral efficacy. Biodistribution analysis of pAAu + pNAu, pAAu-Erl + pNAu, pAAu + pNAu-Dox, and pAAu–Erl + pNAu-Dox (12.4 Au mg kg−1 of mouse, Erl:Dox = 1:1) in MCF-7 (A) and A431 (B) tumor-bearing mice. Infrared thermal images of MCF-7 and A431 tumor-bearing mice before and after NIR laser irradiation (808 nm, 0.5 W cm−2, for 10 min) (C). Analysis of the tumor growth curve in MCF-7 (D) and A431 (E) tumor-bearing mice. * p < 0.05 analysed by ANOVA. Photos and histological analysis of MCF-7 (F,H) and A431 (G,I) tumor-bearing mice. Reprinted with permission from [68]. Copyright (2019) Elsevier.
Figure 2. In vivo evaluation of the pAAu-Erl plus pNAu-Dox antitumoral efficacy. Biodistribution analysis of pAAu + pNAu, pAAu-Erl + pNAu, pAAu + pNAu-Dox, and pAAu–Erl + pNAu-Dox (12.4 Au mg kg−1 of mouse, Erl:Dox = 1:1) in MCF-7 (A) and A431 (B) tumor-bearing mice. Infrared thermal images of MCF-7 and A431 tumor-bearing mice before and after NIR laser irradiation (808 nm, 0.5 W cm−2, for 10 min) (C). Analysis of the tumor growth curve in MCF-7 (D) and A431 (E) tumor-bearing mice. * p < 0.05 analysed by ANOVA. Photos and histological analysis of MCF-7 (F,H) and A431 (G,I) tumor-bearing mice. Reprinted with permission from [68]. Copyright (2019) Elsevier.
Nanomaterials 12 03166 g002
Figure 3. Analysis of the tumor growth curves for 10 days after treatment with iron oxide nanomaterials, mean ± SE and n = 5 (A). TUNEL histological analysis of mice tumors at day 9. Statistical significance is represented by the asterisk (*** p < 0.001). TUNEL assay for mice tumors on day 9 following various treatments (B). Reprinted with permission from [105]. Copyright (2019) Elsevier.
Figure 3. Analysis of the tumor growth curves for 10 days after treatment with iron oxide nanomaterials, mean ± SE and n = 5 (A). TUNEL histological analysis of mice tumors at day 9. Statistical significance is represented by the asterisk (*** p < 0.001). TUNEL assay for mice tumors on day 9 following various treatments (B). Reprinted with permission from [105]. Copyright (2019) Elsevier.
Nanomaterials 12 03166 g003
Figure 4. Photos of the tumor-bearing mice at 1, 3, 7, 10 and 14 days (A) and SPC-A-1 tumors at day 14 (B) after the treatment of the nanomaterials. Tumor growth curves in the control group; 0 µL CuS-PEG in LS, 30 µL CuS-PEG in the RS, 50 µL of CuS-PEG in the LL, and 100 µL CuS-PEG in the RL (C). Tumor growth curves in the NIR—irradiated group; 0 µL NS in the LS, 30 μL NS in the RS, 50 μL NS in the LL, and 100 μL NS CuS-PEG in the RL (D). Tumor growth curves in the CuS-PEG + NIR group, 100 μL NS in the LS, 30 μL CuS-PEG in the RS, 50 μL CuS-PEG in the LL, and 100 μL CuS-PEG in the RL (E). Tumor growth volume curve comparison for 100 μL (30 μg/mL) CuS-PEG +NIR in the RL; 100 μL (30 μg/mL) CuS-PEG in the RL, and 100 μL NS in the RL (F). LS—left shoulder administration; RS—right shoulder administration; LL—left leg administration; RL—right leg administration; NS—normal saline solution. Reprinted with permission from [111]. Copyright (2017) Elsevier.
Figure 4. Photos of the tumor-bearing mice at 1, 3, 7, 10 and 14 days (A) and SPC-A-1 tumors at day 14 (B) after the treatment of the nanomaterials. Tumor growth curves in the control group; 0 µL CuS-PEG in LS, 30 µL CuS-PEG in the RS, 50 µL of CuS-PEG in the LL, and 100 µL CuS-PEG in the RL (C). Tumor growth curves in the NIR—irradiated group; 0 µL NS in the LS, 30 μL NS in the RS, 50 μL NS in the LL, and 100 μL NS CuS-PEG in the RL (D). Tumor growth curves in the CuS-PEG + NIR group, 100 μL NS in the LS, 30 μL CuS-PEG in the RS, 50 μL CuS-PEG in the LL, and 100 μL CuS-PEG in the RL (E). Tumor growth volume curve comparison for 100 μL (30 μg/mL) CuS-PEG +NIR in the RL; 100 μL (30 μg/mL) CuS-PEG in the RL, and 100 μL NS in the RL (F). LS—left shoulder administration; RS—right shoulder administration; LL—left leg administration; RL—right leg administration; NS—normal saline solution. Reprinted with permission from [111]. Copyright (2017) Elsevier.
Nanomaterials 12 03166 g004
Table 1. Gold-based metallic-polymer nanoconjugates/nanohybrids, their physicochemical properties, and therapeutic applications (N.D.—non disclosed, N.A.—not applicable).
Table 1. Gold-based metallic-polymer nanoconjugates/nanohybrids, their physicochemical properties, and therapeutic applications (N.D.—non disclosed, N.A.—not applicable).
MetalMorphologyModificationSize (nm)Surface Charge (mV)LoadingIn VitroIn VivoApplicationRef.
GoldRodsUCST polymer (P(AAm-co-AN)-DDAT), metalloproteinase 2 (MMP-2)-sensitive peptidesLength ≈ 48.04; Width ≈ 12.08N.D.Doxorubicin (DOX)HepG2 cellsHepG2 tumor-bearing micePTT (λex = 808 nm) and chemotherapy[33]
Mesoporous silica; D-α-Tocopherol polyethylene glycol 1000 succinate (TPGS), and Hyaluronic acid (HA)Length ≈ 85; Width ≈ 64−3 ± 5 and −10 ± 4 for TPGS/HA ratios of 1:1 and 4:1, respectivelyN.A.HeLa cellsN.A.PTT (λex = 780 nm)[86]
Mesoporous silica, HA, and polyethyleneimine (PEI)Length: 88 ± 5; Width: 63 ± 5;−10 ± 2Acridine Orange (AO)HeLa cellsN.A.PTT (λex = 750 nm) and chemotherapy[87]
SpheresPoly(ethylene glycol) (PEG) and Lactofferin (LF)5N.D.N.A.Caco-2, U87MG cellsGBM tumor-bearing micePTT (λex = 532 nm)[54]
StarsPolydopamine (PDA) and Folic acid (FA)149 ± 3−19 ± 2.7DOXMCF-7, MCF-7/ADR, NIH/3T3,
and
HaCaT cells
MCF-7/ADR bearing micePTT (λex ≈ 800 nm) and chemotherapy[56]
Dendritic polyglicerol (dPG) and HA68.113.9Retinoic acid (RA)MDA-MB-231 cells4T1 tumor-bearing micePTT (λex ≈ 800 nm) and chemotherapy[88]
PEG and CD133 antibody≈120−22.47IR780/DTXPC3 cellsPC3 tumor-bearing micePTT (λex = 810 nm), PDT, and chemotherapy[89]
CagesPoly (acrylic acid) (pA) or Poly(NIPAM-co-AM) (pN)for pA(Au) ≈130; N.D. for pN(Au)≈−4 for pA(Au) formulation at pH 7.4;
N.D. for pN (Au) formulation
pA(Au)-loaded with Erl and pN(Au) loaded with DOXA431 or MCF-7 cellsA431 or MCF-7 tumor-bearing micePTT (λex ≈ 800 nm for both formulations) and chemotherapy[68]
PVP, PEG, and anti-heat shock protein (HSP) monoclonal antibody61.2 ± 4.85−8.2 ± 1.25N.A.4T14T1 tumor-bearing micePTT (λex ≈ 808 nm)[90]
Abbreviations—PDT: Photodynamic Therapy; PTT: Photothermal Therapy.
Table 2. Iron-based metallic-polymer nanoconjugates/nanohybrids, their physicochemical properties, and therapeutic applications (N.D.—non disclosed, N.A.—not applicable).
Table 2. Iron-based metallic-polymer nanoconjugates/nanohybrids, their physicochemical properties, and therapeutic applications (N.D.—non disclosed, N.A.—not applicable).
MetalMorphologyModificationSize (nm)Surface Charge (mV)LoadingLongitudinal/Transverse Proton RelaxivityIn VitroIn VivoApplicationsRef.
RingGO (graphene oxide) and CREKA (Cys-Arg-Glu-Lys-Ala)223.322 ± 0.4N.A.N.D.4T1 cells4T1 tumor-bearing miceMTD and MTT[6]
IronSpheresHA conjugated with dopamine (HA-DA)60.7−16N.A.r1: 41.3 mM−1A549, HepG2, CT26, B16F10, and 4T1 cells4T1, B16F10, and CT26 tumor-bearing miceMRI[103]
PLGA, silica, Polyaniline (PANI), and R8-RGD20622.8Cisplatinr2: 258.5 mM−1 s−1A549 cellsA549 tumor-bearing micePTT (Strong Absorption in NIR region), MRI, and chemotherapy[104]
PEG, RGD, D-Glucosamine32.31 ± 0.71−30.2 ± 0.76N.A.r2: 554 mM−1 s−14T1 cells4T1 tumor-bearing miceMRI and hyperthermia[105]
PEI, PLGA, and HA159.5 ± 2.3−9.1Olaparib (Olb)Saturation magnetizations: 21.08 emu/gMDA-MB-231 cellsMDA-MB-231 tumor-bearing miceRMF and chemotherapy[106]
PLGA, gold shell, and Herceptin285.7 ± 81.4N.D.DOXr2: 345.31 ± 23.06 mM−1 s−1BT474, MCF, and BT474/Adr cellsBT474 tumor-bearing miceMRI, PTT (λex ≈ 750–800 nm), and chemotherapy[107]
AS1411 and PLGA201.87 ± 1.60−10.67 ± 0.25N.A.N.D.MCF-7 cellsMCF-7 tumor-bearing micePA/US imaging and PTT (λex = 635 nm),[108]
HA-SS-PLA≈11N.D.PTXN.D.HeLa cellsHeLa tumor-bearing miceChemotherapy[109]
SheetsPDA (polydopamine), and rGO (reduced graphene oxide)251−27.5N.A.N.D.MCF-7 cellsN.A.MRI, PTT (Strong Absorption in NIR region), and PDT[110]
Abbreviations—MRI: Magnetic Resonance Imaging; MTD: Magnetothermodynamic therapy; MTT: Magnetothermal Therapy; PA: Photoacoustic Imaging; PDT: Photodynamic Therapy; PTT: Photothermal Therapy; RMF: Rotating Magnetic Field Therapy; US: Ultrasound Imaging.
Table 3. Copper-based metallic-polymer nanoconjugates/nanohybrids, their physicochemical properties, and therapeutic applications (N.D.—non disclosed, N.A.—not applicable).
Table 3. Copper-based metallic-polymer nanoconjugates/nanohybrids, their physicochemical properties, and therapeutic applications (N.D.—non disclosed, N.A.—not applicable).
MetalMorphologyModificationSize (nm)Surface Charge (mV)LoadingIn VitroIn VivoApplicationsRef.
CopperSpheresLanthanide-doped nanoparticles and PEG45N.D.N.A.HeLa cellsCervical cancer tumor xenograftNIR-II luminescence imaging/CT/MRI, CDT, and PDT[121]
p-(OEOMA-co-MEMA)285−17.2TAPPCT26 cellsCT26 tumor-bearing micePA/PI, PTT (Band from visible to NIR), PDT, and SDT[122]
DSPE-PEG modified with Lanreotide186.1 ± 5.2−16.4 ± 0.1DocetaxelPC-3 cellsPC-3 tumor-bearing micePA, PI, PTT (Band between 700 and 1000 nm), PDT, and chemotherapy[123]
PLGA, PDA, and PEG288 (Higher MW-PLGA);
257 (Lower MW-PLGA)
−18.7 (Higher MW-PLGA);
−22.2 (Lower MW-PLGA)
N.A.Cal-33 cellsN.A.MRI, PTT (N.D.), and chemotherapy[124]
HA/PEI330.716.9DisulfiramEca109Eca109 tumor-bearing miceChemotherapy and FL[125]
PEG-NH2 and
PCL-SS-P(DPA/GMA/MP)
151.5 ± 2.2−17.1 ± 1.7DoxL929 and 4T14T1 tumor-bearing micePTT (Strong absorption in the NIR region), and chemotherapy[126]
HA and PDA106−19.43DoxHeLa and 4T14T1 tumor-bearing micePA, PTT (N.D), CDT, and chemotherapy[127]
FrameworkPluronic F127186.4 ± 16.7−1.2 ± 0.1O24T1 and HeLa cells4T1 tumor-bearing micePDT (Band from visibile to NIR)[128]
CubesBSA and PEG-FA60N.A.N.A.HepG2 cellsN.A.PTT (Band from visible to NIR) and chemotherapy[129]
Abbreviations—CT: Computed Tomography; PA: Photoacoustic Imaging; PDT; Photodynamic Therapy; PTT: Photothermal Therapy; SDT: Sonodynamic Therapy.
Table 4. Summary of the Platinum/Silver/Zinc-based metallic-polymer nanoconjugates/nanohybrids, their physicochemical properties, and therapeutic applications (N.D.—non disclosed, N.A.—not applicable).
Table 4. Summary of the Platinum/Silver/Zinc-based metallic-polymer nanoconjugates/nanohybrids, their physicochemical properties, and therapeutic applications (N.D.—non disclosed, N.A.—not applicable).
MetalMorphologyModificationSize (nm)Surface Charge (mV)LoadingIn VitroIn VivoApplicationsRef.
PlatinumSpheresPDA and Folate≈100N.A.Indocyanine Green (ICG)MCF-7Breast cancer tumor xenograftPA, FL, PTT (λex ≈ 700–800 nm), and PDT[136]
PEG120−14.6DOX4T14T1 tumor-bearing miceEDT and chemotherapy[138]
HA38 ± 6−31 ± 1N.A.MDA-MB-231 (CD44+) and PC9 (CD44-)MDA-MB-231 tumor-bearing micePI and PTT (N.D.)[139]
PEG119.7−1.6 ± 0.4Cisplatin and IR7804T14T1 tumor-bearing mice/Hepatocellular Carcinoma Patient Derived XenograftPI, FL, PTT (λex = 780 nm), and chemotherapy[144]
SilverGlobular irregular shapeBSA and PEG131.5 ± 2.7−34.68 ± 0.6ICGB16F10 cellsB16F10 tumor-bearing micePTT (λex ≈ 790 nm)[143]
SpheresPolythiourea and PEG25–30N.D.N.A.A549A549 tumor-bearing miceFL[145]
HA104 ± 6.2−30N.A.4T14T1 tumor-bearing miceFL and RT[146]
DotsFA modified DSPE-PEG2000200−30.84N.A.HeLa and A549 cellsHeLa tumor-bearing miceFL/PA imaging and PTT (Strong absorption in the visible and NIR region)[147]
ZincSpheresPVP40≈5−3.6N.A.SW480 and HEK293T cellsSW480 tumor-bearing micePDT (N.D.)[133]
PDA≈175−21.7DOX and DNAzymeA549 cellsA549 tumor-bearing miceFL, PI, GT, PTT (N.D.), and chemotherapy[148]
PEG and RGD112.0 ± 3.2−14.6 ± 5.2PTX4T1 cells4T1 tumor-bearing miceMRI, NIRFI, and chemotherapy[149]
PEG and PLGAPLGA-ZnNPc-NP = 141;
PLGA-ZnPc-NPs = 152
PLGA-ZnNPc-NPs = 4.8;
PLGA-ZnPc-NPs = 5.1
N.A.MCF-7 cellsDMBA-induced breast cancer-bearing miceFL and PDT (N.D.)[150]
Abbreviations—CDT: Chemodynamic Therapy; EDT: Electrodynamic Therapy; FL: Fluorescence Imaging; GT: Gene Therapy; MRI: Magnetic Resonance Imaging; NIRFI: Near Infrared Fluorescence Imaging; PA: Photoacoustic Imaging; PDT: Photodynamic Therapy; PI: Photothermal Imaging; PTT: Photothermal Therapy; RT: Radiotherapy.
Table 5. Summary of clinical trials comprising metallic-polymer nanoconjugates/nanohybrids (N.A.—not applicable).
Table 5. Summary of clinical trials comprising metallic-polymer nanoconjugates/nanohybrids (N.A.—not applicable).
NameDescriptionApplicationAdministration RouteType of CancerClinical Trials Identifier (Phase)ResultsRef.
Magnablate®Iron oxide nanoparticles Magnetic HyperthermiaIntratumoralProstate CancerNCT02033447 (Early Phase I): CompletedNo results yet available[154]
Nanotherm®Iron oxide nanoparticles Magnetic HyperthermiaIntratumoralBrain tumorApproved by the EMA in 2010[37]
IntratumoralProstate CarcinomaNCT05010759: Still recruiting (Phase not applicable)No results yet availableN.A.
AuroLase®PEGylated silica-gold nanoshell
(AuroShell®)
Laser-activated termal ablationIntravenousMetastic lung tumorsNCT01679470: Phase not applicableNo results yet available[156]
Refractory and/or recurrent head and neck tumorsNCT00848042: Phase not applicableNo results yet availableN.A.
Laser-activated termal ablation combined with MRI/US fusion technology for focal ablationNeoplastic Prostate tissueNCT02680535 and NCT04240639 (extension of the previous): Phase not applicableNo results yet availableN.A.
NU-0129®Spherical gold nanoparticle conjugated with siRNA oligonucleotidesTargeting BCL2L12 oncogeneIntravenousGlioblastoma multiforme or Gliosarcoma TreatmentNCT03020017: CompletedNo results provide about the antitumor efficacy[158]
Abbreviations—EMA: European Medicines Agency; MRI: Magnetic Resonance Imaging; US: Ultrasound Imaging.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Figueiredo, A.Q.; Rodrigues, C.F.; Fernandes, N.; de Melo-Diogo, D.; Correia, I.J.; Moreira, A.F. Metal-Polymer Nanoconjugates Application in Cancer Imaging and Therapy. Nanomaterials 2022, 12, 3166. https://doi.org/10.3390/nano12183166

AMA Style

Figueiredo AQ, Rodrigues CF, Fernandes N, de Melo-Diogo D, Correia IJ, Moreira AF. Metal-Polymer Nanoconjugates Application in Cancer Imaging and Therapy. Nanomaterials. 2022; 12(18):3166. https://doi.org/10.3390/nano12183166

Chicago/Turabian Style

Figueiredo, André Q., Carolina F. Rodrigues, Natanael Fernandes, Duarte de Melo-Diogo, Ilídio J. Correia, and André F. Moreira. 2022. "Metal-Polymer Nanoconjugates Application in Cancer Imaging and Therapy" Nanomaterials 12, no. 18: 3166. https://doi.org/10.3390/nano12183166

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