Skip to content
BY 4.0 license Open Access Published by De Gruyter December 31, 2020

Cobalt magnetic nanoparticles as theranostics: Conceivable or forgettable?

  • Claudia Petrarca EMAIL logo , Anna Maria Poma , Giulia Vecchiotti , Giovanni Bernardini , Qiao Niu , Anna Giulia Cattaneo , Mario Di Gioacchino and Enrico Sabbioni
From the journal Nanotechnology Reviews

Abstract

Superparamagnetic nanoparticles, exposed to an external variable magnetic field, undergo rapid excitation/relaxation. So-called soft magnets, typically iron-based, rapidly and completely relax when the magnetic field returns to zero. Instead, cobalt-based (CoB) hard magnets retain residual magnetization, a characteristic related with the procedure for nanoparticles (NPs) production. Many researchers are still attracted by the potential of CoB NPs for theranostics as multifaced signal probes for imaging, microrobots, enhanced thermo/radiation therapy, and drug release. Since iron oxide NPs are the only magnetic NPs approved for human use, they are of reference for analyzing the potential of the disregarded CoB NPs. In vitro observed toxicity of CoB NPs, largely attributable to cobalt ions and other chemical species released by dissolution, excluded them from further investigations in humans. Nevertheless, experimental evidences documenting the in vivo toxicity of engineered CoB NPs remain very few. The surface functionalization adds newer properties and could improve the biocompatibility of NPs, critical for the clinical exploitation. In our opinion, it would be worth to further exploit the potential of finely tunable properties of CoB NPs in in vivo systems in order to establish a systematic database of properties and effects suitable for human application.

Graphical abstract

1 Introduction

The discovery and exploitation of the superparamagnetism in nanoparticles began with the description of their ferrofluidic behavior [1] which opened the way to the development of nanomaterials with impressive chemical and physical oddities and exponential increase of research on their potential applications in biomedicine. Ferrofluids are colloidal suspensions of very fine (typical size of 10 nm) magnetic particles dispersed in a polar or nonpolar liquid carrier [2]. Application of an external magnetic field to such ferrofluidic materials leads to a very quick increase in magnetization, which is rapidly reduced (or abrogated) when the magnetic field is removed. This behavior is revealed by superparamagnetic iron oxide (FeO)-based nanoprobes that, when excited, acquire a sharp appearance relative to the surrounding environment. For this reason, they attract the interest of researchers as negative contrast enhancing agents for Magnetic Resonance Imaging (MRI) [3,4]. However, these “soft materials” show low residual magnetism (M r) (Figure 1), one main physical property for this type of medical application.

Figure 1 
               Simulated hysteresis curve for superparamagnetic soft (red) and hard (green) NPs. A similar behavior was recorded comparing cobalt-ferrite and magnetite NPs [19]. The saturation magnetization (M
                  s), residual magnetization (M
                  r), and coercivity (Hc) are indicated.
Figure 1

Simulated hysteresis curve for superparamagnetic soft (red) and hard (green) NPs. A similar behavior was recorded comparing cobalt-ferrite and magnetite NPs [19]. The saturation magnetization (M s), residual magnetization (M r), and coercivity (Hc) are indicated.

Moreover, since their early development occurred more than two decades ago, only one carbohydrate-coated iron-based nanoprobe has been approved by FDA for clinical use [5]. Nevertheless, uncoated iron oxide nanoprobes show some degree of toxicity in a therapeutic setting [6]. Furthermore, increased exhaled breath condensate concentrations of lipid, nucleic, and protein oxidation markers were detected in a dose-response correspondence to the environmental monitoring concentrations of NPs, in workers exposed to Fe2O3- and Fe3O4-NPs compared to control individuals [7]. Besides, cobalt and their oxides, when organized at the nanoscale, show a steep increase of the magnetic order and the other related parameters, so that superparamagnetism appears [8]. Differently from iron-based, CoB NPs behave magnetically as “hard materials” and could be worthily used because they own several specific biochemical and physical properties. For instance, in MRI applications they do not generate confounding signals in the presence of unsaturated hemoglobin, as certain Fe-based NPs do [9]. Other distinctive characteristics of CoB NPs (i.e., large magnetic multiaxial anisotropy, high Curie’s temperature, high chemical stability, and mechanical hardness) might purposely be combined with each another to confer to them special “smart” qualities finalized to diagnostic platforms and therapeutic approaches, such as drug release or gene-modifying interventions. Ideally, smart agents should be easy to administer and have low toxicity, short half-life, and fast clearance depending on the diagnostic or therapeutic use. However, also materials with longer half-life should be taken under consideration to allow the follow-up of the evolution of lesions in days and up to months. In addition, magnetic NPs possess juxtaposed superparamagnetic, thermal, photoacoustic, and electric properties [10]. They can also be modified with protective and/or functional (bio)molecules at their surface (preventing dissolution, early phagocytosis, and toxicity), improving the emitted signal (radioactive, bioluminescent, or fluorescent) or target-specific sites (tissues, cells, enzymatic activity, or genome) [11]. Indeed, CoB NPs are attractive under all those regards and would offer great flexibility of applications in a variety of experimental settings. Regrettably, CoB NPs have been addressed by several experimental findings as more instable and toxic than Fe-based NPs, limiting their further exploitation in living beings and humans as a precautious rule. Otherwise, in the present review we critically reevaluate the current knowledge of the crucial physical (matter) parameters governing the synthesis of superparamagnetic of CoB NPs that could be modified/regulated ad hoc to obtain desired outcomes upon interaction with biological systems.

2 The core of superparamagnetic CoB NPs

The magnetic behavior, along with other various outstanding features of metallic NPs, arises from a complex interplay between their chemical nature, lattice symmetry, and shape; furthermore, size appears to be the most relevant parameter, with larger size determining a more disordered surface and longer relaxation time [12] (Figure 2).

Figure 2 
               Magnetic properties of NPs are influenced by various factors, either intrinsic (upper part) or environmental (lower part).
Figure 2

Magnetic properties of NPs are influenced by various factors, either intrinsic (upper part) or environmental (lower part).

As bulk material, cobalt is a ferromagnetic metal. However, when the particle radius is reduced to the nanosize, cobalt behaves as superparamagnetic material, due to the quantum effect [13]. If the radius is small enough, one single NP contains just one magnetic domain that can be finely tuned for theranostic purposes [14]. In turn, the value of the critical radius (r c) for a controllable single magnetic domain is mainly determined by three physical entities: uniaxial anisotropy (K u), vacuum permeability (μ 0), and saturation magnetization (M s) [15], according to:

r c 9 ( A K u ) 1 2 μ 0 M s 2

Table 1 reports the experimental values of residual magnetization (M r) of various different sized CoB NPs with different oxidation state. Zerovalent Co-NPs show the smallest radius (<5 nm), corresponding to the lowest levels of magnetic disorder, conferring better superparamagnetic properties, compared with corresponding Co-oxide NPs [16]. Oxidation reduces the values of saturation magnetization (M s) of Co-NPs, showing the lowest values [17]. The critical size for good ferroparamagnetic properties approximately spans from 5 nm (Co/FePt alloy NPs) to 20 nm (Fe-oxides) [17], CoFe2O4 NPs and iron-based NPs having intermediate M s values [18]. The relatively high energy emitted as heat (thermal energy) allows complete relaxation of soft magnets when the external magnetic field ceases (Figure 3a). Systems with randomly distributed (larger) size or not uniform crystal structure (anisotropic), such as Fe-based NPs, display a “soft magnet” behavior [19]. Instead, CoB NPs behave as “hard magnets,” since they are characterized by uniform size and crystal symmetry [20,21] (Figure 3b). Given the dependence of so many and complex physicochemical properties of NPs, the volume-to-specific surface area ratio, rather than the diameter alone, appears to be the most suitable parameter to describe the phenomenon of superparamagnetism. According to the current consensus about the definition of NP, the critical limit value does not exceed 60 m2 cm−3 [22].

Table 1

Physical-chemistry characteristics and related magnetic parameters of CoB NPs

Magnetic CoB NPs Chemical physical characteristics Magnetic properties Ref.
Size (nm) Shape Crystal symmetry Coercitivity (Hc, Oe) Saturation magnetization (M s,a emu/g) Residual magnetization (M r,a emu/g)
Zerovalent Co@C 123 121 4.1 [140]
Co 40 370 137 [141]
Oxides Co3O4, prepared at 175°C 11 Spherical Cubic 0.137 [23]
Co3O4, prepared at 200°C 0.225
Co3O4, prepared at 250°C 0.325
Ferrites CoFe2O4 25 Tetragonal Cubic, fcc 1,209 64.06 25 [89]
CoFe2O4@AOT (0.05 g) 25 Tetragonal Cubic, fcc 1,880 69.62 33
CoFe2O4@AOT (0.1 g) 25 Tetragonal Cubic, fcc 2,399 71.54 38
CoFe2O4@AOT (0.2 g) 25 Tetragonal Cubic, fcc 2,415 73.78 39
CoFe2O4@AOT (0.5 g) 25 Tetragonal Cubic, fcc 2,550 79.05 43
Fe45Co55 9 Spherical Cubic 43 111.6 [142]
Fe45Co55@graphite <10 Spherical 70 145
CoFe2O4 21 Hexagonal Spinel, cubic 16.3 54.9 15.7 [143]
CoFe2O4@PVP 24 83.2 60.1 26.7
CoFe2O4@PEG 23 84 63 28
CoFe2O4, co-precipitation 33 Spherical Spinel, cubic 1492.6 60.85 29.34 [144]
CoFe2O4, hydrothermal 14 507.7 56.88 21.44
Fe70Co30 15 220 [145]
CoFe2O4 (annealing: 400°C) 1.725 36 [146]
CoFe2O4 (annealing: 850°C) 728 63
CoFe2O4 (annealing: 1,000°C) 298 78
CoFe2O4 25–30 Polyhedral Spinel 5.2 5 [147]
CoFe2O4@PA90% 5.3 6
CoFe2O4@PA80% 5.3 6
CoFe2O4@PA70% 5.3 7
CoFe2O4@PA60% 5.3 8
CoFe2O4@PA60% 5.4 14
CoFe2O4 (300 K) 6 Spherical Spinel, cubic 78 52 [19]
CoFe2O4 (2.5 K) 99 66
CoFe2O4 10 Cubic and hexagonal, mixed 1.656 31 15 [148]
Zn0.5Co0.5Fe2O4 3.9 Spherical 14.02 29.2 68.85 × 10−3 [149]
Dextran-coated Zn0.5Co0.5Fe2O4 12.27 23.6 22.56 × 10−3
CoFe2O4 36 Spherical Inverse spinel, cubic 1.1 69 28 [150]
CoFe2O4@PVA (stirred) 17 1.013 16 [151]
CoFe2O4@PVA (sonicated) 7 170 7
CoFe2O4 9 Single spinel, cubic 234.32 40 [152]

Abbreviations: AOT: the surfactant sodium bis(2-ethylhexyl)sulphosuccinat.

  1. a

    emu/g: equals SI unit (A2/kg).

Figure 3 
               Magnetic relaxation of ferrofluids: stable suspensions of magnetic nanoparticles (MNPs) are ferromagnetic when an external magnetic field applies, and return to their original paramagnetic state, with null residual magnetization, when the external field ceases. Two mechanisms could interact in the magnetic relaxation process: (a) intrinsic rotation of molecules inside the MNPs, which are distributed into a single domain. Chemical nature, volume, and temperature are main regulators of the process that depends on the Nèel relaxation. (b) Extrinsic rotation of MNPs in the colloidal matrix. The driving Brownian rotational motion of single particles depends on shape, temperature, size, and opsonization. More details in ref. [138].
Figure 3

Magnetic relaxation of ferrofluids: stable suspensions of magnetic nanoparticles (MNPs) are ferromagnetic when an external magnetic field applies, and return to their original paramagnetic state, with null residual magnetization, when the external field ceases. Two mechanisms could interact in the magnetic relaxation process: (a) intrinsic rotation of molecules inside the MNPs, which are distributed into a single domain. Chemical nature, volume, and temperature are main regulators of the process that depends on the Nèel relaxation. (b) Extrinsic rotation of MNPs in the colloidal matrix. The driving Brownian rotational motion of single particles depends on shape, temperature, size, and opsonization. More details in ref. [138].

During the demagnetization, heat develops (according to Nèel’s relaxation) (Figure 3), which is the main property of multifunctional NPs to be useful for theranostics.

CoB NPs, more prone to behave as hard magnets than their iron-based counterparts, develop higher and more prolonged temperature rise [23]. Beside size and anisotropy, the magnetization of CoB NPs can be influenced by additional characteristics, such as shape [24] and viscosity of the external matrix [25] (Figure 2).

The method used to synthetize the CoB NPs determines their anisotropy degree and, thereof, the critical size, that greatly varies consequently.

3 Synthesis of magnetic CoB NPs

In order to produce CoB NPs with appealing magnetic properties for medical purposes, the synthesis process should be carried on under controlled conditions. Cobalt can crystalize in a multiplicity of phases (with similar energy levels) and oxidizes rapidly at the surface, where the thin layer of oxide would hamper the completion of the synthesis reaction and uniformity of the final product, thus interfering with magnetic properties [26]. The most relevant methods of CoB NPs synthesis are hereby recalled:

  1. Solvothermal method. It is used for the synthesis of superparamagnetic zerovalent metallic cobalt NPs (CoNPs). The process starts with the induction of a single nucleation, under strictly controlled temperature in a reducing or hypoxic environment. Uncontrolled surface modification (aggregation of new nuclei or oxidation) is avoided by constraining the surface of the growing particles with the addition of a surfactant to the reaction mixture; then, a regular lattice of 2 nm spherical NPs with cubic face-centered symmetry and acting as hard magnets are produced by adding hydrazine and triethanolamine [20] (Figure 4a); 6–9.5 nm NPs with epsilon symmetry are synthesized in the presence of oleic acid:trioctylphosphine oxide 5:1 [27,28] (Figure 4b); 5 nm NPs with hexagonal close-packed symmetry are synthetized in oleic acid:diphenylether 1:1 [29] (Figure 4c).

  2. Single nucleation core. The synthesis of Co-oxide or/Co-ferrite NPs (CoONPs and CoFeNPs) relies on the same principle of single nucleation core followed by controlled growth of the crystal, as summarized in supplementary files (Figures S1 and S2). The precursor could be an inorganic salt or an organic compound, such as cobalt nitrate or cobalt acetate, for cobalt oxide, or salts of iron and cobalt for stoichiometric CoFe NPs. Extremely high temperatures are required for the nucleation and growth of cobalt-ferrite NPs (up to 5,500°C), whereas much lower is needed for cobalt (130–180°C) and cobalt oxide (50–75°C). The composition of the reaction mixture, the reaction time, and temperature are the most important variables to control the final size and the uniformity of the sample [23,30];

  3. Sol–gel method. It utilizes a stable dispersion of micelles that collides, coalesces, and breaks until a uniform precipitate develops in the presence of cobalt nitrate. The addition of appropriate solvent destroys the micelles and produces a final sample of monophasic cobalt oxide NPs after calcination at high temperature [31,32]. Uniform CoFe NPs, smaller than those produced in aqueous solutions, can be obtained through a similar procedure [33]. The experimental comparison of the above-described methods, introducing ad hoc procedural changes (i.e., molecular ratios in the reaction mixture or lower temperature of calcination), led to the production of NPs of the intended size, shape, nanostructure, crystalline symmetry, and magnetic properties, confirming the outmost relevance of rigorous control of the conditions to get such an accomplishment [8].

  4. Biosynthetic process. More recently, CoB NPs are being obtained through procedures as technological response to the quest for environment-friendly synthesis of advanced materials. High superparamagnetic Co–(Fe)-based NPs of stoichiometric (CoFe2O4) and nonstoichiometric composition (i.e., magnetite) with a diameter of 8, 16, and 15 nm, respectively, are produced by the microorganism Geobacter sulfurreducens [34]. Another new and proficient system is based on the use of the empty capsid of Cowpea Mosaic Virus as a nanoshell bioreactor which, filled with 10% in volume cobalt chloride, allows to build up magnetic NPs of irregular structure [35]. More recently, Aspergillus nidulans was found to be suitable for synthesis of spinel cobalt oxide nanoparticles at an average size of 20.29 nm in spherical shape with sulfur-bearing proteins acting as a capping agent for the synthesized nanoparticles [36]. In general, cobaltous nitrate hexahydrate, cobalt(ii) acetate tetrahydrate, cobalt chloride hexahydrate, and cobalt(ii) acetyl acetonate were used as precursors for the green synthesis of Co3O4 NPs [37].

Figure 4 
               Synthesis of zerovalent CoB NPs via solvothermal route. (a) 2 nm MNPs with face-centered cubic symmetry and acting as hard magnets [20]; (b) 6–9.5 nm, epsilon phase [27]; (c) 5 nm, hexagonal close-packed symmetry [139].
Figure 4

Synthesis of zerovalent CoB NPs via solvothermal route. (a) 2 nm MNPs with face-centered cubic symmetry and acting as hard magnets [20]; (b) 6–9.5 nm, epsilon phase [27]; (c) 5 nm, hexagonal close-packed symmetry [139].

3.1 Surface modification

Surface coating or covering with a shell generally increases the biocompatibility and the stability of NPs, protecting them from dissolution and oxidation [38] which are massive processes for CoB NPs in aqueous medium. The chemical nature and the amount of the coating can moderately influence the magnetization, as shown in Table 1.

Organic coating of CoB NPs provides a means to reduce dissolution and facilitate further functionalization with targeting molecules [39]. Biehl and coworkers [40] reviewed the rationale for surface functionalization with polyzwitterionic molecules concluding that, since the method of production can modify the performance as magnetic materials, the functionalized product should be carefully tested in different environments. Gold shell provides a hard protection against dissolution and oxidation which enhances the biocompatibility of CoB NPs. Notably, the shell does not alter per sé the superparamagnetic behavior of NPs, but the number of particles embedded inside the shell seems to be important [41,42] (Figure 5a). Smart materials have been developed with more complex coatings, such as multiple layers of polymeric and soft structures added to a hard golden shell, for the release of drugs on demand. For instance, drug-loaded NPs reacting to an oscillating external magnetic field with a spatial distortion provisionally release the drug when the field ceases [43] (Figure 5b). In addition, particles shaped in the form of small caps (nanowontons) behave as sensitive magnetic and photoacoustic probes: in this case, the protective Au-coating improved the bioavailability without altering the richness of the signal [10] (Figure 5c).

Figure 5 
                  Effect of coating on magnetic properties of cobalt-based NPs. (a) From the left: zerovalent Co, uncoated particle, large particle in a thin golden shell, small particle in a golden shell, Co MNP in a silver shell, Co-oxide MNP. (b) Microcapsules for controlled release of drugs. The inner core, a cobalt-ferrite magnetic nanoparticle sized 5 nm, is embedded in four layers of organic polymers. A thin shell of Au and five organic polymeric layers complete the capsule. If charged with drugs, it reacts to an external magnetic field with distortion of the capsule, permeabilization of layers, and tunable drug release [43]. (c) “Nanowontons,” bare (left) or coated with Au (right), display magnetic and acoustic properties [10].
Figure 5

Effect of coating on magnetic properties of cobalt-based NPs. (a) From the left: zerovalent Co, uncoated particle, large particle in a thin golden shell, small particle in a golden shell, Co MNP in a silver shell, Co-oxide MNP. (b) Microcapsules for controlled release of drugs. The inner core, a cobalt-ferrite magnetic nanoparticle sized 5 nm, is embedded in four layers of organic polymers. A thin shell of Au and five organic polymeric layers complete the capsule. If charged with drugs, it reacts to an external magnetic field with distortion of the capsule, permeabilization of layers, and tunable drug release [43]. (c) “Nanowontons,” bare (left) or coated with Au (right), display magnetic and acoustic properties [10].

More recently, a new material was engineered by cross-linking Co–Fe NPs to a pH-sensitive hydrogel allowing rapid magnetization, possibly given by the average orientation of magnetic polar structures embedded in the gel layer, driven by pH-driven swelling and showing residual weak magnetization after shrinking [44]. Notably, drug-loaded magnetic hydrogels can control the release of drugs in response to a variety of stimuli [45]. An even more tempting coating could be that made of a magnetic material different from that employed in the core. Systems with a ferromagnetic–antiferromagnetic interface show a shift of the magnetization axis, while the juxtaposition of soft and hard magnetic phases generated high-performant, permanent NPs [46,47]. The addition of fluorescent dyes to obtain multiple signals was proved to be efficient in CoB NPs coated with a graphene-like surface of carbon, which did not alter the magnetic properties of the particles and provided a shell easy to be functionalized [48]. Such NPs are a promising probe for in vitro assays, with fluorescent signal and recovery on external magnet. Instead, a shell of silica does not preserve the magnetic signal of Co–Fe NPs functionalized with multiple fluorescent dyes [49]. The polymeric system made up with CoO-NPs coated with phosphonomethyl iminodiacetic acid had excellent stability in aqueous medium and, tested in vitro, this material showed anticancer activity without adverse effect on normal cells [50].

The complexity of this matter, describing the changes of magnetic order that follow coating and surface functionalization, as well as size, shape, composition, and methods for preparation, stresses the importance of carefully characterizing newer particles, before considering the exploitation of final products.

4 Cobalt-based NPs as potential theranostic agents

Nowadays, iron-based NPs are the sole approved for clinical use by the international drug authorizing agencies and, therefore, the preferred ones for in vivo applications [4,51,52]. Nevertheless, the ability of MNPs to transduce external magnetic field energy into a mechanical or thermal response can be exploited for biomedical applications, with multifaced working hypothesis: in vivo imaging, targeted release of drugs, magnetically driven navigation with delivery of thermo/radio/chemotherapy, as well as long-term follow-up of localized lesions, pathogens, and parasites. In spite of this enormous potential, only 14 human studies with magnetic nanoparticles, all iron-based NPs, are retrieved from the Clinical Trials database at this moment [4,53,54] directed to improve imaging of cancer and treatment of cardiovascular, demyelinating, and inflammatory nervous diseases (Figure 6 and Tables S1–S4).

Figure 6 
               The Cobalt nanoparticles can be toxic at multiple levels, through direct or dissolution-mediated mechanisms. MNPs and dissolved cobalt ions can covalently bound the heme group (hematologic toxicity), stimulate the immunologic system, enhance oxidative damage to the DNA, or injure the cardiovascular system.
Figure 6

The Cobalt nanoparticles can be toxic at multiple levels, through direct or dissolution-mediated mechanisms. MNPs and dissolved cobalt ions can covalently bound the heme group (hematologic toxicity), stimulate the immunologic system, enhance oxidative damage to the DNA, or injure the cardiovascular system.

In the meantime, human studies are disallowed for CoB NPs, whose persistent neglect is consequent of their supposed high instability and toxicity for the human organisms, at multiple levels (Figure 6).

Nevertheless, experimental studies are currently carried on to the aim of improving diagnostics and therapeutic tools by implementing CoB NPs (Figure 7) as exposed in the next paragraphs.

Figure 7 
               The distribution of studies with MNPs deposited in the Clinical Trials. The most numerous are for enhanced imaging of cancer (n = 18), 10 studies are aimed to the treatment and imaging of pathologies in the cardiovascular system, five to the imaging of demyelinating or inflammatory diseases of central nervous system, and five (not shown) to other conditions.
Figure 7

The distribution of studies with MNPs deposited in the Clinical Trials. The most numerous are for enhanced imaging of cancer (n = 18), 10 studies are aimed to the treatment and imaging of pathologies in the cardiovascular system, five to the imaging of demyelinating or inflammatory diseases of central nervous system, and five (not shown) to other conditions.

4.1 Diagnostic imaging: providing improved and multiple signal

Although Fe-based NPs are well-established and widely accepted since long time as paramagnetic probes for clinical applications, CoB NPs, of comparable paramagnetic properties and even more ductile and appealing as hard magnets, have been relegated to design/develop research studies in vitro and in laboratory animals, after toxicity side effects were accidentally described in those systems.

In fact, CoB NPs have been under investigation for the possibility of obtaining multiple signals through, for instance, the functionalization with fluorescent dyes. In addition to the superparamagnetic signal, high sensitivity emission signals were detected making them improved imaging tools [10,49,55]. In a study conducted in rats, the superparamagnetic signal emitted from CoB NPs could be further boosted by the luminescent and radioactive ones by functionalizing the NPs with Ga68 and luciferase [56]; in experimental biology, rhodamine-coated fluorescent cobalt–ferrite magnetic NPs have been used to monitor intracellular miR124a during neuronal differentiation of murine P19 cells [56].

4.2 Magnetic driving and navigation of NPs through the circulation

In a in vivo rodent model, cells loaded with Co–Fe NPs (enclosed in a silica shell or viral capsid) were able to migrate from the site of injection to distant tissues under the guidance of an externally applied magnetic field and accumulating nearby [57]. However, this cell-based approach might show limits in larger organisms where length of migration and depth of accumulation are increased [58]. When a more targeted experiment was performed, magnetic CoB NPs injected in the tail vein of rats accumulated inside the anterior chamber of the eye, where the external magnetic field was applied. Therefore, the particles could certainly navigate a long way through the vascular bed, but, also in this case, the site of accumulation was superficial and the size of the animal very small [59]. In 2011, a significant step was made in the therapeutic field of tumor targeted with the introduction of magnetic NPs as drug-containing cargos which could be guided by an external magnetic field. For the first time, the principles of microfluidics, robotics, and magnetic driving successfully cooperated to selectively address chemotherapy at a deep tumor site, through the circulatory system in living animals. The Magnetic Resonance Navigation (MRN) technique consisted of directional magnetic driving of a biocompatible fluid of 500 µm polymeric capsules filled with Co–Fe magnetic NPs manoeuvred through the vascular system by an external magnetic field [60].

After that, magnetically navigable agents for chemotherapy were investigated on phantom models mimicking liver arteries [61] and circulation [62]. As expected, the most probable detrimental effect would be embolism; furthermore, the ferromagnetic NPs microrobot navigation, under the control of magnetic force, would induce fluidity microvariations without apparent toxicity of conjugated ferromagnetic NPs as such.

4.3 Magnetic hyperthermia and drug release for tumor ablation

For decades, implants of macroseeds of iron alloys have been safely and efficiently applied for magnetically induced ablative thermotherapy of prostate tumors [63]. In phase I clinical trial on thermotherapy of recurrent prostate cancer, the employed iron oxide magnetic NPs showed no side effects [64], as well as the magnetic iron oxide NPs used for intratumoral thermotherapy were described as safe in glioblastoma patients [65]. However, those approaches could not achieve optimal modulation of the heating and induced massive tissue necrosis and incomplete ablation, in relation to dimension and depth of the tumor. Those disadvantages have been overcome by making biocompatible magnetic NPs which, in clinical trials, were shown to safely accumulate deeply in tumor lesions at the effective dose [66] and, when loaded inside stem cells or polymeric microcarriers, efficiently destroy atherosclerotic plaques [67]. Interestingly, the amount of cobalt in magnetic CoFe NPs determines their heating response to a variable magnetic field, in experimental conditions [51,68,69]; in this way, it could be tuned to obtain theranostics suitable for diverse clinical applications [19,70] by varying the quantity of Co purposely. At the cellular level, CoB NPs, like others, enter various cell types through endocytosis [71,73] or diffusion across the plasma membrane [7477] and are retained even when an alternate magnetic field is applied [71]. Remarkably, the endocellular dynamics and fate of CoB NPs are rather complex and are highly relevant to the biological outcome [78] (Figure 8).

Figure 8 
                  Cobalt oxide MNPs in the extracellular space can (A) interact with albumin or histidine, and be confined to the extracellular environment, or enter the cells through endocytosis (B) or passive diffusion (C). Partial dissolution occurs in both spaces, extra- and endocellular, with release of divalent ions of cobalt (D). The divalent ions of cobalt enter the cells through specific receptors and metalloproteins (a). Inside the cell, both MNPs and ions distribute in the cytoplasm and in the nucleus. Both ions and MNPs enhance the production of ROS and the subsequent oxidative stress, with mitochondrial damage and apoptosis (1), lipid and protein peroxidation, with cell membrane damage (2 and 3), and damage to the DNA (4). Note that zerovalent Cobalt NPs do not cross the cell membrane [77,78].
Figure 8

Cobalt oxide MNPs in the extracellular space can (A) interact with albumin or histidine, and be confined to the extracellular environment, or enter the cells through endocytosis (B) or passive diffusion (C). Partial dissolution occurs in both spaces, extra- and endocellular, with release of divalent ions of cobalt (D). The divalent ions of cobalt enter the cells through specific receptors and metalloproteins (a). Inside the cell, both MNPs and ions distribute in the cytoplasm and in the nucleus. Both ions and MNPs enhance the production of ROS and the subsequent oxidative stress, with mitochondrial damage and apoptosis (1), lipid and protein peroxidation, with cell membrane damage (2 and 3), and damage to the DNA (4). Note that zerovalent Cobalt NPs do not cross the cell membrane [77,78].

Furthermore, the viscosity of the milieu in which magnetic NPs are dispersed affects the physical phenomenon driving heat response and needs to be taken into account. For instance, when dispersed into glycerol whose viscosity is similar to that of the endocellular milieu, the thermal effect of NPs is mainly due to magnetic heating mechanism of Néel relaxation [79]. Hence, the best performing NPs would likely be soft magnetic materials (maghemite, magnetite, and iron–platinum). Anyway, also hard magnets such as CoB NPs, expected to develop heat through the Brown relaxation mechanism (Brownian friction), have shown to be suitable for hyperthermia applications in several, but methodologically disparate, experimental studies [80,81,82]. These considerations suggest that thermal properties of magnetic nanoparticles might be conveniently modulated by modifying their own viscosity in the endocellular environment.

Dissolution of CoB NPs is a phenomenon known to take place either outside and inside the cell, releasing potentially dangerous ions of cobalt [78]. Instead, Fe-based NPs, under a constant magnetic field, either in cellular [83] or acellular [84] model systems, form large agglomerates and show low dissolution rate. However, until now no study has been addressed to determine whether CoB NPs release ions under similar conditions of magnetization, relaxation, and heat release. Paradoxically, the ion-mediated cytotoxicity of magnetic CoB NPs might even be a looked-for effect for treating tumor cells.

Magnetic CoB NPs have been found to induce thermal necrosis/apoptosis of melanoma [85], human breast [86], and rat gliosarcoma [87] cancer cells in vitro. Interestingly, human breast cancer exposed to the magnetic CoB NPs shows higher intracellular reactive oxygen species (ROS), found to mediate Co-ions-induced cytotoxicity [73].

However, these few experimental findings, although suggestive, are not enough to support the medical application of CoB NPs [88]. Nevertheless, it is reasonable to forecast the development of innovative CoB NPs by tuning the chemical–physical properties and their operational environment according to the up-to-date findings. This method is useful to overcome problems of toxicity for thermoablative cancer treatment, for the future improvement of implantable or injectable devices, for magnetic navigation, heating, and resonance [64].

4.4 Other potential exploitation of CoB NPs in oncology

Despite the existing interdiction to their clinical use that disincentives the exploitation of their full potential, there are various hints that CoB NPs deserve to be further investigated to produce smart NPs usable in theranostics. Their magnetic properties, tunable and regulated by external magnetic fields, can modulate multiple functions, such as imaging, photothermal, and pharmaceutical targeting of lesions, with minimally invasive procedures. The persistence of particles at the site of lesion allows long-term monitoring of relapse. In particular, surface functionalization with binding/reacting molecules [89], graphite [90], or carbon nanotubes [91] does not alter their magnetic behavior and allows efficient loading and temperature-controlled release of antitumor drugs [92]. Furthermore, functionalization of magnetic NPs with radioactive isotopes, currently extensively investigated in preclinical studies, might produce tools for (potential) simultaneous radiotherapy and hyperthermia and imaging [93].

The displayed toxicity of CoB NPs appears not only related to inherent characteristics, but, at various extents, to the cellular targets. Pancreatic cancer cells are less sensitive than ovarian cancer cells [94]; normal lymphocytes and squamous oral cells, but not lymphoma and oral carcinoma cells, are protected from CoB NPs-related toxicity by coating with organic compounds [50]; and melanoma cells magnetic effectively die by heating only when cobalt is incorporated in the core of ferrite NPs [85]. Normal cells can be exempted from unwanted cobalt toxicity related to therapeutic treatment by the functionalization of CoB NPs with antigens or peptides that selectively bind ligands exclusively expressed by tumors cells. In this way, CoB NPs target and accumulate more efficiently at the site of lesion and inside the target cells [50,95]. Analogously, magnetic targeting, intracellular labelling, and selective removal of ovarian cancer cells from the ascitic fluid in vitro and in vivo are feasible using an external capture magnet [96].

Differently from other metal NPs that need to be PEG-coated to permeate the cell membrane, naked CoB NPs could load a sufficient dose of conjugated doxorubicin drug in the target tumor cell [97].

The conjugation of CoB NPs with porphyrin derivatives produces a powerful, multifunctional therapeutic platform for boron neutron capture, phototherapy, and fluorescence imaging. All these products can accumulate rapidly and in large amounts (80%) in human lung adenocarcinoma A549 cells and their cytoplasmic organelles, resulting to be cytotoxic only when purposely activated with photons or neutrons irradiation, but not in basal conditions [98]. Interestingly, an in vivo study with rats showed that recently developed Co-based nanoclusters could accrue in ovarian cancer tumors and effectively elevate intratumoral temperature (following a single intravenous injection), being nontoxic [99].

Hence, several findings gathered on CoB NPs converge towards its uneven intrinsic noxiousness, under certain usage condition. An ordered database on the fundamentals for knowledge, prediction of potential uses, and commercial restrictions of CoB NPs, still lacking, could prompt further research and development of these magnetic NPs for nanomedicine.

5 Mechanisms of CoB NPs toxicity

Weakening of interest and skepticism arouse on CoB NPs’ application in nanomedicine for their toxicity observed in animal and human tumor cells in vitro [73,100,101,102,103,104] and in vivo [95,105,106]. Despite that, it is important to underline that the intrinsic material characteristics of CoB NPs are not the only ones playing a role in cytotoxicity.

These events result by interaction with the milieu (biochemical and physical factors) where NPs have to operate. Keeping this in mind, those concerns might result in overlooking leading to too negatively conditioned technological advancement. Indeed, theoretical reasoning based on the “Ostwald Ripening phenomenon” predicts that high specific surface area of CoB NPs increases the potential for cobalt ions (Co2+) to be released from these NPs. However, Co2+ release occurs when CoB NPs are surrounded by aqueous milieu. In that case, spontaneous morphological modifications would occur based on the thermodynamically driven atomic exchange between atoms in solution and atoms in the NPs [107]. Then, in water the extent of the dissolution depends on the size and speciation of Co constituting the NPs. It has been shown that both metallic zerovalent Co0- and Co3O4NPs release Co2+ ions in biological media. However, the Co0 form releases greater amount of Co2+, compared to Co-oxide [77,78,108]. Relevantly for the present dissertation, in the extracellular environment the serum proteins of cell culture media appear to bind the released Co2+, subtracting them to the exchange medium-particle and so facilitating the dissolution of Co-particles. Moreover, CoB NPs can be uptaken by cells and internalized inside the organelles [72,78] where they can dissolve into Co2+, which are massively detected in nuclei and mitochondria [78], otherwise present at ultratrace level under physiological conditions.

It is not clear to which degree the toxicity of CoB NPs results from the released Co2+ or to the integer NPs [109]. However, a study with mouse fibroblasts Balb/3T3 found out that Co2+ are statistically more cytotoxic, compared to the corresponding nano- and micro-particulated Co. In addition, it was observed that the dissolution rate was size-dependent, being higher for the nanosized compared to the microsized Co-particles. Remarkably, the cytotoxicity increases according to the increase of the total intracellular Co content, rather than Co exposure concentrations [78]. Moreover, micro and nano forms of zerovalent Co-particles activate toxicologically relevant transcriptional pathways implicated in carcinogenesis and inflammation [110]. Noteworthy, for such toxicological experiments high concentrations of Co species were used, in any case far beyond those achivable for therapeutic aims. Therefore, the extent of intracellular Co accrual and the understanding of the size-dependent dissolution of CoB NPs are the crucial points for addressing their safe use in nanomedicine, by providing protection from dissolution for long-term stability.

Given the described findings on the crucial role of released Co2+ in the mechanisms of CoB NPs-induced toxicity, the molecular basis of Co2+-induced toxic effects must be considered. General mechanisms extensively reviewed by experts in this field [111115] include: (i) formation of ROS in the presence of hydrogen peroxide with generation of hydroxyl radicals (HO˙) via Fenton reaction and irreversible damage by protein, lipid, carbohydrate, and DNA oxidation [114]; (ii) direct interaction, particularly with sulfhydryl groups, and binding to cellular proteins of the redox system (enzymes catalase, heme oxygenase, superoxide dismutase, peroxidase), metabolism (arginase), molecular transport (transferrin, hemoglobin), motility (lymphocyte cytosolic protein), and signalling (phosphodiesterase 3A). This leads to a direct induction of oxidation and loss of biological function [114]. Moreover, Co2+ alter Ca2+ influx into cells [116], acting as a blocker of inorganic calcium channels, and modify glucose metabolism [117]; (iii) displacement of other essential divalent metal ions in the ion center of metal-activated enzymes (i.e., Zn of alkaline phosphatase [118], Zn2+ and Mg2+ of cell proteins [78], Fe2+ of dioxygenases [119]).

Systemic toxic effects are induced when Co/Co2+ ions enter the blood and the lymphatic system and traslocate into different organs. Urinary excretion is the main mechanism of inorganic cobalt clearance from the human body [120], decreasing with time after exposure [111]. Below, we remind some toxicity mechanisms for Co/Co2+-induced systemic effects.

5.1 Hematological effects

Cobalt therapy in the treatment of anemia induces polycythemia and hypothyroidism [121]. Co2+ inhibit the synthesis of heme in vivo by acting on the biosynthesis of 5-delta-amino levulinate and its conversion into heme, resulting in the generation of cobalt protoporphyrin [122]. Heme oxidation in many tissues is also stimulated. In addition, cobalt acts to increase erythropoietin release from damaged renal cells, which stimulates the production of red blood cells [123].

5.2 Thyrotoxic effects

Cobalt impaired thyroid activity and goiter formation [124]. The inhibition of iodine uptake and tyrosine iodinase activity by cobalt prevent the incorporation of iodine into thyroxine [114], being considered the mechanistic basis of the observed hypothyroidism induced by cobalt therapy [125].

5.3 Myocardial effects

Cobalt-induced cardiomyopathy was observed in beer drinkers [126]. Mitochondrial changes are representative of disturbance in energy production/utilization. The irreversible chelation of alpha-lipoic acid –SH groups in the citric acid cycle by Co2+ under hypoxic conditions is considered the mechanistic basis for the pathogenesis of Co-induced cardiotoxicity [127].

5.4 Neurotoxic effects

High levels of cobalt released from metal prosthesis can induce neurotoxic effects involving several neurologically related organ systems (i.e., auditory, ocular, central, and peripheral nervous systems) [128]. Such effects of cobalt are mediated through depletion of neurotransmitters such as dopamine, noradrenaline, and serotonin and presynaptic blockage of calcium channels [128].

5.5 Immunological effects

Metallic cobalt, or in the form of water-soluble ionized salts, is allergenic causing immediate type I as well as delayed type IV hypersensitivity reactions [113], rhinitis, and asthma. IgA and IgE antibodies specific to cobalt have been observed in humans [129]. Cobalt ions can act as hapten binding to serum proteins to form hapten-like complexes to produce immunogenic products that can account for allergic reactions [130]. T-lymphocytes regulate cobalt sensitivity, while cobalt reduces the proliferation in vitro of B and T lymphocytes and the release of IL-2, IL-6, and INF-γ [74]. The release of Co-ions is associated with proinflammatory reaction involving monocyte and lymphocyte reactions [131].

5.6 Carcinogenicity and genotoxicity

Cobalt sulfate and other soluble cobalt(ii) salts are classified as “possibly carcinogenic to humans” (Group 2B), while the mixture dust cobalt/tungsten carbide (Co/WC) as “probably carcinogenic to humans” (group 2A) [132]. Soluble cobalt induces mutagenic effects in mammalian cells in vitro by a direct damage to DNA by a Fenton-like mechanism; or indirectly by a mechanism involving inhibition of repair of DNA damage, particularly concerning the incision and polymerization steps, through interaction with zinc-finger DNA repair proteins, considered the most relevant mechanism for metal carcinogenicity than binding to DNA [133]. Two mechanisms are involved in the Co-particles-induced mutagenic effects: production of ROS by both Fenton and non-Fenton (i.e., in the absence of H2O2) mechanisms, resulting in DNA damage, and release of Co2+ which inhibit DNA repair processes [132].

These above-reported Co2+-related biochemical mechanisms, general or specific for Co2+-induced systemic effects, are relevant also for Co-particles species. In fact, the toxic reaction that produces ROS leads in parallel to the oxidation of metallic cobalt and hence to the formation of ionic species. However, the described cobalt-related systemic effects are in general uncommon because of a high toxicity threshold.

Systemic Co toxicity manifests with a variable presentation of neurological, cardiovascular, and endocrine symptoms, depending on the systemic Co levels (blood/urine). These systemic effects, called “arthroprosthetic cobaltism” syndrome, were initially observed in metal-on-metal implant recipient subjects, accompanying the presence, in the periprosthetic tissue, of electron dense nanosized debris inside the macrophages [113,134,135].

6 Conclusions

The theoretical basis for further research on this nano-substance that might reasonably be voided of toxic features and, therefore suitable in theranostics, is exposed.

Cobalt-based NPs, which behave as hard magnets, could be considered the possible counterpart of the soft magnets composed of iron oxide MNPs. The residual magnetization is higher in MNPs with a core of cobalt oxide or cobalt–ferrite, in comparison with zerovalent cobalt, while selected variants in the procedure for production, such as the temperature or the relative content of iron in cobalt–ferrite MNPs, can tune the saturation magnetization. The crystal symmetry of the final product mainly supports this effect. The surface functionalization with hard or soft shells protects from dissolution and could also change the magnetic properties.

The alternative option of using CoB NPs to overcome some limits of the iron-based soft magnets requires appropriate precautions. It is mandatory for the surface functionalization or shell of CoB NPs to reduce the feared cobalt toxicity since cobalt particles applied in theranostics medicine could result in exposure to ions that may potentially reach the threshold of high toxicity. In addition, cobalt nanotoxicity must be further investigated from a mechanistic point of view. In this context, the aspect of speciation will have to be the core of future studies in this area. In any case, the intracellular distribution of Co2+ released from CoB NPs entering the cells plays a fundamental role in determining the toxic effects of Co-particles. Thus, drawing any conclusions about the differential toxicity of particles relative to ions should be done with great care, as uptake effect relationships must be considered [78,136].

At the present, there is a lack of tiered strategy for the characterization of CoB NPs, which reduces the potential of these tools. Indeed, a range of CoB NPs nanomaterials may offer a powerful platform for biological implementation in theranostics, as improved probes and catalysts. In nanomedicine, the supposed toxicity of Co dictates the choice of iron oxide MNPs, considered more biocompatible. Consequenlty, CoB NPs are poorly exploited in this field [137], with a great waste of potentiality. Only the diagnostic utilizations improve, without any advance in therapy: neither trials nor experimental studies have been performed focusing on thermotherapy with CoB MNPs in medicine, clinical or veterinary, or on other possible therapeutic applications. Nevertheless, under this point of view, CoB MNPs offer several advantages in comparison with those of iron oxide. Their higher residual magnetization allows a more persistent signal; the magnetic heating is more tuneable and, therefore, more compatible with the necessary avoidance of local necrosis. These NPs that do not produce interfering signals with the iron bound to heme could be used as radiation isotope against tumor growth.

In our opinion, it is time to revise the band of these tools in medicine. Even the evaluation of the risk for toxicity could be obsolete, considering newer ways to prevent the dissolution of toxic species of Co. For example, a golden shell may be able to make CoNPs biocompatible, preventing dissolution and modulating magnetism and magnetic heating. Phase I clinical trials could be set up and conclusively determine the fate of these promising, but until now neglected, MNPs, following a severe control of all steps of production, complete characterization of the product, and tired test for toxicity, according to the current approved protocols.

  1. Conflict of interest: The authors declare no conflict of interest regarding the publication of this paper.

References

[1] Papell SS, Faber OC. On the influence of nonuniform magnetic fields on ferromagnetic colloidal sols. NASA technical note; 1968.Search in Google Scholar

[2] Scherer C, Figueiredo Neto AM. Ferrofluids: properties and applications. Braz J Phys. 2005;35:718–27.10.1590/S0103-97332005000400018Search in Google Scholar

[3] Zheng J, Ren W, Chen T, Jin Y, Li A, Yan K, et al. Recent advances in superparamagnetic iron oxide based nanoprobes as multifunctional theranostic agents for breast cancer imaging and therapy. Curr Med Chem. 2017;25:3001–16.10.2174/0929867324666170705144642Search in Google Scholar

[4] Arias LS, Pessan JP, Vieira APM, de Lima TMT, Delbem ACB, Monteiro DR. Iron oxide nanoparticles for biomedical applications: a perspective on synthesis, drugs, antimicrobial activity, and toxicity. Antibiotics (Basel, Switzerland). 2018;7:46. 10.3390/antibiotics7020046.Search in Google Scholar

[5] Shah A, Mankus CI, Vermilya AM, Soheilian F, Clogston JD, Dobrovolskaia MA. Feraheme® suppresses immune function of human T lymphocytes through mitochondrial damage and mitoROS production. Toxicol Appl Pharmacol. 2018;350:52–63.10.1016/j.taap.2018.04.028Search in Google Scholar

[6] Landry R, Jacobs PM, Davis R, Shenouda M, Bolton WK. Pharmacokinetic study of ferumoxytol: a new iron replacement therapy in normal subjects and hemodialysis patients. Am J Nephrol. 2005;25:400–10.10.1159/000087212Search in Google Scholar

[7] Pelclova D, Zdimal V, Fenclova Z, Vlckova S, Turci F, Corazzari I, et al. Markers of oxidative damage of nucleic acids and proteins among workers exposed to TiO2(nano) particles. Occup Env Med. 2016;73:110–8.10.1136/oemed-2015-103161Search in Google Scholar

[8] Luisetto I, Pepe F, Bemporad E. Preparation and characterization of nano cobalt oxide. J Nanopart Res. 2008;10:59–67.10.1007/s11051-008-9365-4Search in Google Scholar

[9] Bjørnerud A, Briley-Sæbø K, Johansson LO, Kellar KE. Effect of NC100150 injection on the1H NMR linewidth of human whole blood ex vivo: dependency on blood oxygen tension. Magn Reson Med. 2000;44:803–7.10.1002/1522-2594(200011)44:5<803::AID-MRM19>3.0.CO;2-KSearch in Google Scholar

[10] Bouchard LS, Anwar MS, Liu GL, Hann B, Xie ZH, Gray JW, et al. Picomolar sensitivity MRI and photoacoustic imaging of cobalt nanoparticles. Proc Natl Acad Sci USA. 2009;106:4085–9.10.1073/pnas.0813019106Search in Google Scholar

[11] Wu K, Su D, Liu J, Saha R, Wang JP. Magnetic nanoparticles in nanomedicine: a review of recent advances. Nanotechnology. 2019;30:502003.10.1088/1361-6528/ab4241Search in Google Scholar

[12] Kolhatkar AG, Jamison AC, Litvinov D, Willson RC, Lee TR. Tuning the magnetic properties of nanoparticles. Int J Mol Sci. 2013;14:15977–6009.10.3390/ijms140815977Search in Google Scholar PubMed PubMed Central

[13] Qiao R, Zhang XL, Qiu R, Li Y, Kang YS. Fabrication of superparamagnetic cobalt nanoparticles-embedded block copolymer microcapsules. J Phys Chem C. 2007;111:2426–9.10.1021/jp067534sSearch in Google Scholar

[14] Zarschler K, Rocks L, Licciardello N, Boselli L, Polo E, Garcia KP, et al. Ultrasmall inorganic nanoparticles: state-of-the-art and perspectives for biomedical applications. Nanomedicine. 2016;12:1663–701.10.1016/j.nano.2016.02.019Search in Google Scholar PubMed

[15] Bedanta S, Barman A, Kleemann W, Petracic O, Seki TJ. Magnetic nanoparticles: a subject for both fundamental research and applications. Nanomaterials. 2013;2013:952540.10.1155/2013/952540Search in Google Scholar

[16] Tucek J, Sofer Z, Bouša D, Pumera M, Holá K, Malá A, et al. Retracted article: air-stable superparamagnetic metal nanoparticles entrapped in graphene oxide matrix. Nat Commun. 2016;7:12879.10.1038/ncomms12879Search in Google Scholar PubMed PubMed Central

[17] Amiri S, Shokrollahi H. The role of cobalt ferrite magnetic nanoparticles in medical science. Mater Sci Eng C Mater Biol Appl. 2013;33:1–8.10.1016/j.msec.2012.09.003Search in Google Scholar PubMed

[18] Liu F, Laurent S, Roch A, Vander Elst L, Muller RN. Size-controlled synthesis of CoFe2O4 nanoparticles potential contrast agent for MRI and investigation on their size-dependent magnetic properties. J Nanomater. 2013;2013:462540.10.1155/2013/462540Search in Google Scholar

[19] Psimadas D, Baldi G, Ravagli C, Comes Franchini M, Locatelli E, Innocenti C, et al. Comparison of the magnetic, radiolabeling, hyperthermic and biodistribution properties of hybrid nanoparticles bearing CoFe2O4 and Fe3O4 metal cores. Nanotechnology. 2014;25:025101.10.1088/0957-4484/25/2/025101Search in Google Scholar PubMed

[20] Alagiri M, Muthamizhchelvan C, Hamid SBA. Synthesis of superparamagnetic cobalt nanoparticles through solvothermal process. J Mater Sci Mater Electron. 2013;24:4157–60.10.1007/s10854-013-1375-zSearch in Google Scholar

[21] Yang JY, Kim JY, Lee JS, Woo SJ, Kwak JS, Hong JP, et al. Inverted hysteresis loops observed in a randomly distributed cobalt nanoparticle system. Phys Rev B Condens Matter Mater Phys. 2008;78:094415.10.1103/PhysRevB.78.094415Search in Google Scholar

[22] Roebben G, Rauscher H, Amenta V, Aschberger K, Sanfeliu A, Calzolai L, et al. Towards a review of the EC recommendation for a definition of the term “nanomaterial” part 2: assessment of collected information concerning the experience with the definition; 2014.Search in Google Scholar

[23] Farhadi S, Safabakhsh J, Zaringhadam P. Synthesis, characterization, and investigation of optical and magnetic properties of cobalt oxide (Co3O4) nanoparticles. J Nanostruct Chem. 2013;3:69.10.1186/2193-8865-3-69Search in Google Scholar

[24] Jamet M, Wernsdorfer W, Thirion C, Mailly D, Dupuis V, Mélinon P, et al. Magnetic anisotropy of a single cobalt nanocluster. Phys Rev Lett. 2001;86:4676.10.1103/PhysRevLett.86.4676Search in Google Scholar PubMed

[25] Skumryev V, Stoyanov S, Zhang Y, Hadjipanayis G, Givord D, Nogués J. Beating the superparamagnetic limit with exchange bias. Nature. 2003;423:850–3.10.1038/nature01687Search in Google Scholar PubMed

[26] Bautin VA, Seferyan AG, Nesmeyanov MS, Usov NA. Magnetic properties of polycrystalline cobalt nanoparticles. AIP Adv. 2017;7:045103.10.1063/1.4979889Search in Google Scholar

[27] Samia ACS, Hyzer K, Schlueter JA, Qin CJ, Jiang JS, Bader SD, et al. Ligand effect on the growth and the digestion of Co nanocrystals. J Am Chem Soc. 2005;127:4126–7.10.1021/ja044419rSearch in Google Scholar PubMed

[28] Fuller RO, Goh B-M, Koutsantonis GA, Loedolff MJ, Saunders M, Woodward RC. A simple procedure for the production of large ferromagnetic cobalt nanoparticles. Dalton Trans. 2016;45:11983–9.10.1039/C6DT01935HSearch in Google Scholar PubMed

[29] Wang ZY, Rao H, Deng MF, Fan YT, Hou HW. Photocatalytic H2 generation based on noble-metal-free binuclear cobalt complexes using visible-light. Phys Chem Chem Phys. 2013;15:16665.10.1039/c3cp51550hSearch in Google Scholar PubMed

[30] Zhou L, Xu J, Li X, Wang F. Metal oxide nanoparticles from inorganic sources via a simple and general method. Mater Chem Phys. 2006;97:137–42.10.1016/j.matchemphys.2005.07.062Search in Google Scholar

[31] Co AC, Liu J, Serebrennikova I, Abel CM, Birss VI. Structural and electrochemical studies of Co oxide films formed by the sol–gel technique. J Mater Sci. 2005;40:4039–52.10.1007/s10853-005-1994-8Search in Google Scholar

[32] Shokrollahi H. Structure, synthetic methods, magnetic properties and biomedical applications of ferrofluids. Mater Sci Eng C. 2013;33:2476–87.10.1016/j.msec.2013.03.028Search in Google Scholar PubMed

[33] Georgiadou V, Kokotidou C, Le Droumaguet B, Carbonnier B, Choli-Papadopoulou T, Dendrinou-Samara C. Oleylamine as a beneficial agent for the synthesis of CoFe2O4 nanoparticles with potential biomedical uses. J Chem Soc Dalt Trans. 2014;43:6377–88.10.1039/C3DT53179ASearch in Google Scholar

[34] Coker VS, Telling ND, Van Der Laan G, Pattrick RAD, Pearce CI, Arenholz E, et al. Harnessing the extracellular bacterial production of nanoscale cobalt ferrite with exploitable magnetic properties. ACS Nano. 2009;3:1922–8.10.1021/nn900293dSearch in Google Scholar PubMed

[35] Jaafar M, Aljabali AAA, Berlanga I, Mas-Ballesté R, Saxena P, Warren S, et al. Structural insights into magnetic clusters grown inside virus capsids. ACS Appl Mater Interf. 2014;6:20936–42.10.1021/am505682xSearch in Google Scholar PubMed

[36] Vijayanandan AS, Balakrishnan RM. Photostability and electrical and magnetic properties of cobalt oxide nanoparticles through biological mechanism of endophytic fungus Aspergillus nidulans. Appl Phys A Mater Sci Process. 2020;126:1.10.1007/s00339-020-3395-xSearch in Google Scholar

[37] David SA, Veeraputhiran V, Vedhi C. Biosynthesis of cobalt oxide nanoparticles – a short review. J Nanosci Technol. 2019;5:734–7.10.30799/jnst.S01.19050308Search in Google Scholar

[38] Dobbrow C, Schmidt AM. Improvement of the oxidation stability of cobalt nanoparticles. Beilstein J Nanotechnol. 2012;3:75–81.10.3762/bjnano.3.9Search in Google Scholar PubMed PubMed Central

[39] Hegde V, Srivastava C. Synergetic effect of size and morphology of cobalt ferrite nanoparticles on proton relaxivity. IET Nanobiotechnology. 2014;8:184–9.10.1049/iet-nbt.2013.0009Search in Google Scholar PubMed

[40] Biehl P, von der Lühe M, Dutz S, Schacher F. Synthesis, characterization, and applications of magnetic nanoparticles featuring polyzwitterionic coatings. Polymers (Basel). 2018;10:91.10.3390/polym10010091Search in Google Scholar PubMed PubMed Central

[41] Ruano M, Martínez L, Huttel Y. Investigation of the working parameters of a single magnetron of a multiple ion cluster source: determination of the relative influence of the parameters on the size and density of nanoparticles. Dataset Pap Sci. 2013;2013:597023.10.1155/2013/597023Search in Google Scholar

[42] Hrubovčák P, Zeleňáková A, Zeleňák V, Kovác J. Superparamagnetism in cobalt nanoparticles coated by protective gold layer. Acta Phys Pol A Pol Acad Sci. 2014;126:216–7.10.12693/APhysPolA.126.216Search in Google Scholar

[43] Lu Z, Prouty MD, Quo Z, Golub VO, Kumar CSSR, Lvov YM. Magnetic switch of permeability for polyelectrolyte microcapsules embedded with Co@Au nanoparticles. Langmuir. 2005;21:2042–50.10.1021/la047629qSearch in Google Scholar PubMed

[44] Van Berkum S, Biewenga PD, Verkleij SP, Van Zon JBA, Boere KWM, Pal A, et al. Swelling enhanced remanent magnetization of hydrogels cross-linked with magnetic nanoparticles. Langmuir. 2015;31:442–50.10.1021/la503390wSearch in Google Scholar PubMed

[45] Il Kim J, Lee BS, Chun CJ, Cho JK, Kim SY, Song SC. Long-term theranostic hydrogel system for solid tumors. Biomaterials. 2012;33:2251–9.10.1016/j.biomaterials.2011.11.083Search in Google Scholar PubMed

[46] Zeng H, Li J, Liu JP, Wang ZL, Sun S. Exchange-coupled nanocomposite magnets by nanoparticle self-assembly. Nature. 2002;420:395–8.10.1038/nature01208Search in Google Scholar PubMed

[47] Inderhees SE, Borchers JA, Green KS, Kim MS, Sun K, Strycker GL, et al. Manipulating the magnetic structure of Co core/CoO shell nanoparticles: implications for controlling the exchange bias. Phys Rev Lett. 2008;101:11702.10.1103/PhysRevLett.101.117202Search in Google Scholar PubMed

[48] Kainz QM, Reiser O. polymer- and dendrimer-coated magnetic nanoparticles as versatile supports for catalysts, scavengers, and reagents. Acc Chem Res. 2014;47:667–77.10.1021/ar400236ySearch in Google Scholar PubMed

[49] Sung CK, Hong KA, Lin S, Lee Y, Cha J, Lee JK, et al. Dual-modal nanoprobes for imaging of mesenchymal stem cell transplant by MRI and fluorescence imaging. Korean J Radiol. 2009;10:613.10.3348/kjr.2009.10.6.613Search in Google Scholar PubMed PubMed Central

[50] Chattopadhyay S, Dash SK, Ghosh T, Das S, Tripathy S, Mandal D, et al. Anticancer and immunostimulatory role of encapsulated tumor antigen containing cobalt oxide nanoparticles. J Biol Inorg Chem. 2013;18:957–73.10.1007/s00775-013-1044-ySearch in Google Scholar PubMed

[51] Wang C, Hsu CH, Li Z, Hwang LP, Lin YC, Chou PT, et al. Effective heating of magnetic nanoparticle aggregates for in vivo nano-theranostic hyperthermia. Int J Nanomed. 2017;12:6273–87.10.2147/IJN.S141072Search in Google Scholar PubMed PubMed Central

[52] Cattaneo AG, Gornati R, Sabbioni E, Chiriva-Internati M, Cobos E, Jenkins MR, et al. Nanotechnology and human health: risks and benefits. J Appl Toxicol. 2010;30:730–44.10.1002/jat.1609Search in Google Scholar PubMed

[53] Weissig V, Guzman-Villanueva D. Nanopharmaceuticals (part 2): products in the pipeline. Int J Nanomed. 2015;10:1245.10.2147/IJN.S65526Search in Google Scholar PubMed PubMed Central

[54] Nardecchia S, Sánchez-Moreno P, de Vicente J, Marchal JA, Boulaiz H. Clinical trials of thermosensitive nanomaterials: an overview. Nanomaterials. 2019;9:191.10.3390/nano9020191Search in Google Scholar PubMed PubMed Central

[55] Alam SR, Shah ASV, Richards J, Lang NN, Barnes G, Joshi N, et al. Ultrasmall superparamagnetic particles of iron oxide in patients with acute myocardial infarction: early clinical experience. Circ Cardiovasc Imaging. 2012;5:559–65.10.1161/CIRCIMAGING.112.974907Search in Google Scholar PubMed

[56] Hwang DW, Song IC, Lee DS, Kim S. Smart magnetic fluorescent nanoparticle imaging probes to monitor microRNAs. Small. 2010;6:81–8.10.1002/smll.200901262Search in Google Scholar PubMed

[57] Ye H, Shen Z, Yu L, Wei M, Li Y. Manipulating nanoparticle transport within blood flow through external forces: an exemplar of mechanics in nanomedicine. Proc R Soc A Math Phys Eng Sci. 2018;474:20170845.10.1098/rspa.2017.0845Search in Google Scholar PubMed PubMed Central

[58] Weber W, Lienhart C, Daoud-El Baba M, Grass RN, Kohler T, Müller R, et al. Magnet-guided transduction of mammalian cells and mice using engineered magnetic lentiviral particles. J Biotechnol. 2009;141:118–22.10.1016/j.jbiotec.2009.02.023Search in Google Scholar PubMed

[59] Dengler M, Saatchi K, Dailey JP, Matsubara J, Mikelberg FS, Häfeli UO, et al. Targeted delivery of magnetic cobalt nanoparticles to the eye following systemic administration. AIP Conf. Proc. 2010;1311:329–36.10.1063/1.3530034Search in Google Scholar

[60] Pouponneau P, Leroux J-C, Soulez G, Gaboury L, Martel S. Co-encapsulation of magnetic nanoparticles and doxorubicin into biodegradable microcarriers for deep tissue targeting by vascular MRI navigation. Biomaterials. 2011;32:3481–6.10.1016/j.biomaterials.2010.12.059Search in Google Scholar PubMed

[61] Pouponneau P, Soulez G, Beaudoin G, Leroux J-C, Martel S. MR imaging of therapeutic magnetic microcarriers guided by magnetic resonance navigation for targeted liver chemoembolization. Cardiovasc Intervent Radiol. 2014;37:784–90.10.1007/s00270-013-0770-4Search in Google Scholar PubMed

[62] Folio D, Ferreira A. Two-dimensional robust magnetic resonance navigation of a ferromagnetic microrobot using pareto optimality. IEEE Trans Robot. 2017;33:583–93.10.1109/TRO.2016.2638446Search in Google Scholar

[63] Deger S, Taymoorian K, Boehmer D, Schink T, Roigas J, Wille AH, et al. Thermoradiotherapy using interstitial self-regulating thermoseeds: an intermediate analysis of a phase II trial. Eur Urol. 2004;45:574–80.10.1016/j.eururo.2003.11.012Search in Google Scholar PubMed

[64] Johannsen M, Gneveckow U, Thiesen B, Taymoorian K, Cho CH, Waldöfner N, et al. Thermotherapy of prostate cancer using magnetic nanoparticles: feasibility, imaging, and three-dimensional temperature distribution. Eur Urol. 2007;52:1653–62.10.1016/j.eururo.2006.11.023Search in Google Scholar PubMed

[65] Maier-Hauff K, Ulrich F, Nestler D, Niehoff H, Wust P, Thiesen B, et al. Efficacy and safety of intratumoral thermotherapy using magnetic iron-oxide nanoparticles combined with external beam radiotherapy on patients with recurrent glioblastoma multiforme. J Neurooncol. 2011;103:317–24.10.1007/s11060-010-0389-0Search in Google Scholar PubMed PubMed Central

[66] Hainfeld JF, Huang HS. Intravenous magnetic nanoparticle cancer hyperthermia. Int J Nanomed. 2013;8:2521.10.2147/IJN.S43770Search in Google Scholar PubMed PubMed Central

[67] Kharlamov AN, Tyurnina AE, Veselova VS, Kovtun OP, Shur VY, Gabinsky JL. Silica–gold nanoparticles for atheroprotective management of plaques: results of the NANOM-FIM trial. Nanoscale. 2015;7:8003–15.10.1039/C5NR01050KSearch in Google Scholar

[68] Céspedes MV, Unzueta U, Tatkiewicz W, Sánchez-Chardi A, Conchillo-Solé O, Álamo P, et al. In vivo architectonic stability of fullyde novodesigned protein-only nanoparticles. ACS Nano. 2014;8:4166–76.10.1021/nn4055732Search in Google Scholar PubMed

[69] Kekalo K, Shubitidze F, Meyers R, Yaqub R, Baker I. Magnetic heating of Fe-Co ferrites: experiments and modeling. Nano Life. 2016;6:1650007.10.1142/S1793984416500070Search in Google Scholar PubMed PubMed Central

[70] Iatridi Z, Vamvakidis K, Tsougos I, Vassiou K, Dendrinou-Samara C, Bokias G. Multifunctional polymeric platform of magnetic ferrite colloidal superparticles for luminescence, imaging, and hyperthermia applications. ACS Appl Mater Interf. 2016;8:35059–70.10.1021/acsami.6b13161Search in Google Scholar PubMed

[71] Bregar VB, Lojk J, Šuštar V, Veranič P, Pavlin M. Visualization of internalization of functionalized cobalt ferrite nanoparticles and their intracellular fate. Int J Nanomed. 2013;8:919.10.2147/IJN.S38749Search in Google Scholar PubMed PubMed Central

[72] Ortega R, Bresson C, Darolles C, Gautier C, Roudeau S, Perrin L, et al. Low-solubility particles and a Trojan-horse type mechanism of toxicity: the case of cobalt oxide on human lung cells. Part Fibre Toxicol. 2014;11:1.10.1186/1743-8977-11-14Search in Google Scholar PubMed PubMed Central

[73] Sabbioni E, Fortaner S, Farina M, Del Torchio R, Olivato I, Petrarca C, et al. Cytotoxicity and morphological transforming potential of cobalt nanoparticles, microparticles and ions in Balb/3T3 mouse fibroblasts: an in vitro model. Nanotoxicology. 2014;8:455–64.10.3109/17435390.2013.796538Search in Google Scholar PubMed

[74] Castellani ML, Salini V, Frydas S, Donelan J, Madhappan B, Petrarca C, et al. Interleukin-31: a new cytokine involved in inflammation of the skin. Int J Immunopathol Pharmacol. 2006;19:11.10.1177/205873920601900101Search in Google Scholar

[75] Papis E, Rossi F, Raspanti M, Dalle-Donne I, Colombo G, Milzani A, et al. Engineered cobalt oxide nanoparticles readily enter cells. Toxicol Lett. 2009;189:253–9.10.1016/j.toxlet.2009.06.851Search in Google Scholar PubMed

[76] Petrarca C, Clemente E, Di Giampaolo L, Mariani-Costantini R, Leopold K, Schindl R, et al. Palladium nanoparticles induce disturbances in cell cycle entry and progression of peripheral blood mononuclear cells: paramount role of ions. J Immunol Res. 2014;2014:295092–8.10.1155/2014/295092Search in Google Scholar PubMed PubMed Central

[77] Bossi E, Zanella D, Gornati R, Bernardini G. Cobalt oxide nanoparticles can enter inside the cells by crossing plasma membranes. Sci Rep. 2016;6:1.10.1038/srep22254Search in Google Scholar PubMed PubMed Central

[78] Sabbioni E, Fortaner S, Farina M, Del Torchio R, Petrarca C, Bernardini G, et al. Interaction with culture medium components, cellular uptake and intracellular distribution of cobalt nanoparticles, microparticles and ions in Balb/3T3 mouse fibroblasts. Nanotoxicology. 2014;8:88–99.10.3109/17435390.2012.752051Search in Google Scholar PubMed

[79] Weaver JB, Kuehlert E. Measurement of magnetic nanoparticle relaxation time: measurement of magnetic nanoparticle relaxation time. Med Phys. 2012;39:2765–70.10.1118/1.3701775Search in Google Scholar PubMed PubMed Central

[80] Fortin JP, Gazeau F, Wilhelm C. Intracellular heating of living cells through Néel relaxation of magnetic nanoparticles. Eur Biophys J. 2008;37:223–8.10.1007/s00249-007-0197-4Search in Google Scholar PubMed

[81] Habib AH, Ondeck CL, Chaudhary P, Bockstaller MR, McHenry ME. Evaluation of iron-cobalt/ferrite core-shell nanoparticles for cancer thermotherapy. J Appl Phys. 2008;103:07A307.10.1063/1.2830975Search in Google Scholar

[82] Soukup D, Moise S, Céspedes E, Dobson J, Telling ND. In situ measurement of magnetization relaxation of internalized nanoparticles in live cells. ACS Nano. 2015;9:231–40.10.1021/nn503888jSearch in Google Scholar PubMed

[83] Khlusov IA, Zagrebin LV, Shestov SS, Itin VI, Sedoi VS, Feduschak TA, et al. Colony-forming activity of unipotent hemopoietic precursors under the effect of nanosized ferrites in a constant magnetic field in vitro. Bull Exp Biol Med. 2008;145:151–7.10.1007/s10517-008-0026-5Search in Google Scholar PubMed

[84] Tschulik K, Compton RG. Nanoparticle impacts reveal magnetic field induced agglomeration and reduced dissolution rates. Phys Chem Chem Phys. 2014;16:13909–13.10.1039/C4CP01618ASearch in Google Scholar PubMed

[85] Fantechi E, Innocenti C, Zanardelli M, Fittipaldi M, Falvo E, Carbo M, et al. A Smart platform for hyperthermia application in cancer treatment: cobalt-doped ferrite nanoparticles mineralized in human ferritin cages. ACS Nano. 2014;8:4705–19.10.1021/nn500454nSearch in Google Scholar PubMed

[86] Bohara RA, Thorat ND, Chaurasia AK, Pawar SH. Cancer cell extinction through a magnetic fluid hyperthermia treatment produced by superparamagnetic Co–Zn ferrite nanoparticles. RSC Adv. 2015;5:47225–34.10.1039/C5RA04553CSearch in Google Scholar

[87] Hoque SM, Huang Y, Cocco E, Maritim S, Santin AD, Shapiro EM, et al. Improved specific loss power on cancer cells by hyperthermia and MRI contrast of hydrophilic FexCo1-xFe2O4 nanoensembles: higher SLP and T2 relaxivity of hydrophilic nanoensembles. Contrast Media Mol Imaging. 2016;11:514–26.10.1002/cmmi.1713Search in Google Scholar PubMed

[88] Suriyanto EYKNG, Kumar SD. Physical mechanism and modeling of heat generation and transfer in magnetic fluid hyperthermia through Néelian and Brownian relaxation: a review. Biomed Eng Online. 2017;16:36.10.1186/s12938-017-0327-xSearch in Google Scholar PubMed PubMed Central

[89] Zhang Y, Fei C, Liu Y, Wang R, Yan G, Xiong R, et al. The effect of surface modification on the magnetic properties of CoFe2O4 nano-particles synthesized by the hydrothermal method. J Nanosci Nanotechnol. 2010;10:6395–9.10.1166/jnn.2010.2518Search in Google Scholar PubMed

[90] Sherlock SP, Tabakman SM, Xie L, Dai H. Photothermally enhanced drug delivery by ultrasmall multifunctional FeCo/graphitic shell nanocrystals. ACS Nano. 2011;5:1505–12.10.1021/nn103415xSearch in Google Scholar PubMed PubMed Central

[91] Sun C, Liu Y, Ding W, Gou Y, Xu K, Xia Ding GQ. Synthesis and characterization of superparamagnetic CoFe2O4/MWCNT hybrids for tumor-targeted therapy. J Nanosci Nanotechnol. 2013;13:236–41.10.1166/jnn.2013.6711Search in Google Scholar PubMed

[92] Ho BN, Pfeffer CM, Singh ATK. Update on nanotechnology-based drug delivery systems in cancer treatment. Anticancer Res. 2017;37:5975.Search in Google Scholar

[93] Zhang L, Chen H, Wang L, Liu T, Yeh J, Lu G, et al. Delivery of therapeutic radioisotopes using nanoparticle platforms: potential benefit in systemic radiation therapy. Nanotechnol Sci Appl. 2010;3:159.Search in Google Scholar

[94] Pašukonienė V, Mlynska A, Steponkienė S, Poderys V, Matulionytė M, Karabanovas V, et al. Accumulation and biological effects of cobalt ferrite nanoparticles in human pancreatic and ovarian cancer cells. Medicina (Kaunas). 2014;50:237–44.10.1016/j.medici.2014.09.009Search in Google Scholar PubMed

[95] Chattopadhyay S, Dash SK, Tripathy S, Das B, Mandal D, Pramanik P, et al. Toxicity of cobalt oxide nanoparticles to normal cells; an in vitro and in vivo study. Chem Biol Interact. 2015;226:58–71.10.1016/j.cbi.2014.11.016Search in Google Scholar PubMed

[96] Scarberry KE, Dickerson EB, McDonald JF, Zhang ZJ. Magnetic nanoparticle − peptide conjugates for in vitro and in vivo targeting and extraction of cancer cells. J Am Chem Soc. 2008;130:10258–62.10.1021/ja801969bSearch in Google Scholar PubMed

[97] Javed KR, Ahmad M, Ali S, Butt MZ, Nafees M, Butt AR, et al. Comparison of doxorubicin anticancer drug loading on different metal oxide nanoparticles. Medicine (USA). 2015;94:e617.10.1097/MD.0000000000000617Search in Google Scholar PubMed PubMed Central

[98] Efremenko AV, Ignatova AA, Grin MA, Sivaev IB, Mironov AF, Bregadze VI, et al. Chlorin e6fused with a cobalt-bis(dicarbollide) nanoparticle provides efficient boron delivery and photoinduced cytotoxicity in cancer cells. Photochem Photobiol Sci. 2014;13:92–102.10.1039/C3PP50226KSearch in Google Scholar

[99] Albarqi HA, Wong LH, Schumann C, Sabei FY, Korzun T, Li X, et al. Biocompatible nanoclusters with high heating efficiency for systemically delivered magnetic hyperthermia. ACS Nano. 2019;13:6383–95.10.1021/acsnano.8b06542Search in Google Scholar PubMed PubMed Central

[100] Abudayyak M, Gurkaynak TA, Özhan G. In vitro evaluation of cobalt oxide nanoparticle-induced toxicity. Toxicol Ind Health. 2017;33:646–54.10.1177/0748233717706633Search in Google Scholar PubMed

[101] Akhtar MJ, Ahamed M, Alhadlaq HA, Alshamsan A. Nanotoxicity of cobalt induced by oxidant generation and glutathione depletion in MCF-7 cells. Toxicol In Vitro. 2017;40:94–101.10.1016/j.tiv.2016.12.012Search in Google Scholar PubMed

[102] Alarifi S, Ali D, Suliman AO, Ahamed M, Siddiqui MA, Al-Khedhair AA. Oxidative stress contributes to cobalt oxide nanoparticles-induced cytotoxicity and DNA damage in human hepatocarcinoma cells. Int J Nanomed. 2013;8:189.10.2147/IJN.S37924Search in Google Scholar PubMed PubMed Central

[103] Horev-Azaria L, Kirkpatrick CJ, Korenstein R, Marche PN, Maimon O, Ponti J, et al. Predictive toxicology of cobalt nanoparticles and ions: comparative in vitro study of different cellular models using methods of knowledge discovery from data. Toxicol Sci. 2011;122:489–501.10.1093/toxsci/kfr124Search in Google Scholar PubMed

[104] Kwon YM, Xia Z, Glyn-Jones S, Beard D, Gill HS, Murray DW. Dose-dependent cytotoxicity of clinically relevant cobalt nanoparticles and ions on macrophages in vitro. Biomed Mater. 2009;4:025018. 10.1088/1748-6041/4/2/025018.Search in Google Scholar PubMed

[105] Ghoreishi SM, Najafzadeh H, Mohammadian B, Rahimi E, Afzalzadeh MR, Kazemi Varnamkhasti M, et al. Effect of cobalt nano-particles on serum biochemical and histopathological changes in liver and kidney of lambs. Iran J Vet Sci Technol. 2013;5:1.Search in Google Scholar

[106] Novak S, Drobne D, Golobič M, Zupanc J, Romih T, Gianoncelli A, et al. Cellular internalization of dissolved cobalt ions from ingested CoFe2O4 nanoparticles: in vivo experimental evidence. Env Sci Technol. 2013;47:5400–8.10.1021/es305132gSearch in Google Scholar PubMed

[107] Casals E, Pfaller T, Duschl A, Oostingh GJ, Puntes VF. Hardening of the nanoparticle-protein corona in metal (Au, Ag) and oxide (Fe3O4, CoO, and CeO2) nanoparticles. Small. 2011;7:3479–86.10.1002/smll.201101511Search in Google Scholar PubMed

[108] Gornati R, Pedretti E, Rossi F, Cappellini F, Zanella M, Olivato I, et al. Zerovalent Fe, Co and Ni nanoparticle toxicity evaluated on SKOV‐3 and U87 cell lines. J Appl Toxicol. 2016;36:385–93.10.1002/jat.3220Search in Google Scholar PubMed PubMed Central

[109] Colognato R, Bonelli A, Ponti J, Farina M, Bergamaschi E, Sabbioni E, et al. Comparative genotoxicity of cobalt nanoparticles and ions on human peripheral leukocytes in vitro. Mutagenesis. 2008;23:377–82.10.1093/mutage/gen024Search in Google Scholar PubMed

[110] Perconti S, Aceto GMM, Verginelli F, Napolitano F, Petrarca C, Bernardini G, et al. Distinctive gene expression profiles in Balb/3T3 cells exposed to low dose cobalt nanoparticles, microparticles and ions: potential nanotoxicological relevance. J Biol Regul Homeost Agents. 2013;27:443.Search in Google Scholar

[111] Simonsen LO, Harbak H, Bennekou P. Cobalt metabolism and toxicology – a brief update. Sci Total Env. 2012;432:210–5.10.1016/j.scitotenv.2012.06.009Search in Google Scholar PubMed

[112] Paustenbach DJ, Tvermoes BE, Unice KM, Finley BL, Kerger BD. A review of the health hazards posed by cobalt. Crit Rev Toxicol. 2013;43:316–62.10.3109/10408444.2013.779633Search in Google Scholar PubMed

[113] Czarnek K, Terpilowska S, Siwicki AK. Review paper selected aspects of the action of cobalt ions in the human body. Cent Eur J Immunol. 2015;40:236–42.10.5114/ceji.2015.52837Search in Google Scholar PubMed PubMed Central

[114] Cheung AC, Banerjee S, Cherian JJ, Wong F, Butany J, Gilbert C, et al. Systemic cobalt toxicity from total hip arthroplasties: review of a rare condition Part 1 – history, mechanism, measurements, and pathophysiology. Bone Jt J. 2016;98-B:6–13.10.1302/0301-620X.98B1.36374Search in Google Scholar PubMed

[115] Leyssens L, Vinck B, Van Der Straeten C, Wuyts F, Maes L. Cobalt toxicity in humans – a review of the potential sources and systemic health effects. Toxicology. 2017;387:43–56.10.1016/j.tox.2017.05.015Search in Google Scholar PubMed

[116] Henquin JC, Lambert AE. Cobalt inhibition of insulin secretion and calcium uptake by isolated rat islets. Am J Physiol. 1975;228:1669–77.10.1152/ajplegacy.1975.228.6.1669Search in Google Scholar PubMed

[117] Ybarra J, Behrooz A, Gabriel A, Koseoglu MH, Ismail-Beigi F. Glycemia-lowering effect of cobalt chloride in the diabetic rat: increased GLUT1 mRNA expression. Mol Cell Endocrinol. 1997;133:151–60.10.1016/S0303-7207(97)00162-7Search in Google Scholar

[118] Sabbioni E, Girardi F, Marafante E. Replacement of metal in metalloenzymes. A lead-alkaline phosphatase. Biochemistry. 1976;15:271–6.10.1021/bi00647a005Search in Google Scholar

[119] Epstein ACR, Gleadle JM, McNeill LA, Hewitson KS, O'Rourke J, Mole DR, et al. C. elegans EGL-9 and mammalian homologs define a family of dioxygenases that regulate HIF by prolyl hydroxylation. Cell. 2001;107:43–54.10.1016/S0092-8674(01)00507-4Search in Google Scholar

[120] Leggett RW. The biokinetics of inorganic cobalt in the human body. Sci Total Env. 2008;389:259–69.10.1016/j.scitotenv.2007.08.054Search in Google Scholar

[121] Bianchi C, Bertanza C, Mistura L, Pietra R, Sabbioni E. Cobalt-induced hypothyroidism, cardiomyopathy, polycythemia, and hypertricosis in an infant. J Trace Elem Exp Med. 1989;2:311.Search in Google Scholar

[122] Nakamura M, Yasukochi Y, Minakami S. Effect of cobalt on heme biosynthesis in rat liver and spleen1. J Biochem. 1975;78:373–80.10.1093/oxfordjournals.jbchem.a130917Search in Google Scholar

[123] Alexander CS. Cobalt-beer cardiomyopathy. Am J Med. 1972;53:395–417.10.1016/0002-9343(72)90136-2Search in Google Scholar

[124] Nordberg G. Assessment of risks in occupational cobalt exposures. Sci Total Env. 1994;150:201–7.10.1016/0048-9697(94)90151-1Search in Google Scholar

[125] Prager D, Sembrot JT, Southard M. Cobalt-60 therapy of hodgkin’s disease and the subsequent development of hypothyroidism. Cancer. 1972;29:458–60.10.1002/1097-0142(197202)29:2<458::AID-CNCR2820290232>3.0.CO;2-NSearch in Google Scholar

[126] Packer M. Unbelievable folly of clinical trials in heart failure. Circ Hear Fail. 2016;9:4.10.1161/CIRCHEARTFAILURE.116.002837Search in Google Scholar

[127] Webb M. The biological action of cobalt and other metals III. Chelation of cations by dihydrolipoic acid. Biochim Biophys Acta. 1962;65:47–65.10.1016/0006-3002(62)90148-8Search in Google Scholar

[128] Apostoli P, Catalani S, Zaghini A, Mariotti A, Poliani PL, Vielmi V, et al. High doses of cobalt induce optic and auditory neuropathy. Exp Toxicol Pathol. 2013;65:719–27.10.1016/j.etp.2012.09.006Search in Google Scholar

[129] Cirla AM. Cobalt-related asthma: clinical and immunological aspects. Sci Total Env. 1994;150:85–94.10.1016/0048-9697(94)90133-3Search in Google Scholar

[130] Di Gioacchino M, Petrarca C, Lazzarin F, Di Giampaolo L, Sabbioni E, Boscolo P, et al. Immunotoxicity of nanoparticles. Int J Immunopathol Pharmacol. 2011;24:65S–71S.Search in Google Scholar

[131] Caicedo MS, Pennekamp PH, McAllister K, Jacobs JJ, Hallab NJ. Soluble ions more than particulate cobalt-alloy implant debris induce monocyte costimulatory molecule expression and release of proinflammatory cytokines critical to metal-induced lymphocyte reactivity. J Biomed Mater Res A. 2010;93:1312.10.1002/jbm.a.32627Search in Google Scholar PubMed

[132] IARC working group on the evaluation of carcinogenic risks to humans. IARC Monogr Eval Carcinog Risks Hum. 2006;86:1.Search in Google Scholar

[133] Beyersmann D, Hartwig A. Carcinogenic metal compounds: recent insight into molecular and cellular mechanisms. Arch Toxicol. 2008;82:493–512.10.1007/s00204-008-0313-ySearch in Google Scholar PubMed

[134] Mao X, Wong A, Crawford R. Cobalt toxicity – an emerging clinical problem in patients with metal-on-metal hip prostheses. Med J Aust. 2011;194:649–51.10.5694/j.1326-5377.2011.tb03151.xSearch in Google Scholar PubMed

[135] Scharf B, Clement CC, Zolla V, Perino G, Yan B, Elci SG, et al. Molecular analysis of chromium and cobalt-related toxicity. Sci Rep. 2014;4:4. 10.1038/srep05729.Search in Google Scholar PubMed PubMed Central

[136] Hartung T, Sabbioni E. Alternative in vitro assays in nanomaterial toxicology. Wiley Interdiscip Rev Nanomed Nanobiotechnol. 2011;3:545–73.10.1002/wnan.153Search in Google Scholar PubMed

[137] Kudr J, Haddad Y, Richtera L, Heger Z, Cernak M, Adam V, et al. Magnetic nanoparticles: from design and synthesis to real world applications. Nanomaterials. 2017;7:243. 10.3390/nano7090243.Search in Google Scholar PubMed PubMed Central

[138] Rosensweig RE. Magnetic fluids. Annu Rev Fluid Mech. 1987;19:437–61.10.1146/annurev.fl.19.010187.002253Search in Google Scholar

[139] Rao KS, Balaji T, Lingappaa Y, Reddy MRP, Prakash TL. Synthesis and characterisation of ferromagnetic hexagonal cobalt nanoparticles. J Exp Nanosci. 2013;8:162–70.10.1080/17458080.2011.563323Search in Google Scholar

[140] Li H, Li Y, Zhang Y, Liang C, Wang H, Li B, et al. Fabrication and properties of carbon-encapsulated cobalt nanoparticles over NaCl by CVD. Nanoscale Res Lett. 2016;11:11. 10.1186/s11671-016-1645-9.Search in Google Scholar PubMed PubMed Central

[141] Khort A, Podbolotov K, Serrano-García R, Gun’Ko Y. One-step solution combustion synthesis of cobalt nanopowder in air atmosphere: the fuel effect. Inorg Chem. 2018;57:1464–73.10.1021/acs.inorgchem.7b02848Search in Google Scholar PubMed

[142] Tzitzios V, Basina G, Niarchos D, Li W, Hadjipanayis G. Synthesis of air stable FeCo nanoparticles. J Appl Phys. 2011;109:07A313.10.1063/1.3540387Search in Google Scholar

[143] Covaliu CI, Jitaru I, Paraschiv G, Vasile E, Biriş SŞ, Diamandescu L, et al. Core–shell hybrid nanomaterials based on CoFe2O4 particles coated with PVP or PEG biopolymers for applications in biomedicine. Powder Technol. 2013;237:415–26.10.1016/j.powtec.2012.12.037Search in Google Scholar

[144] Khorrami SA, Manuchehri QS. Magnetic properties of cobalt ferrite synthesized by hydrothermal and co-precipitation methods: a comparative study. J Appl Chem Res. 2013;7:15–23.Search in Google Scholar

[145] Wang W, Jing Y, He S, Wang JP, Zhai JP. Surface modification and bioconjugation of FeCo magnetic nanoparticles with proteins. Colloids Surf B Biointerf. 2014;117:449–56.10.1016/j.colsurfb.2013.11.050Search in Google Scholar PubMed

[146] Sedlacik M, Pavlinek V, Peer P, Filip P. Tailoring the magnetic properties and magnetorheological behavior of spinel nanocrystalline cobalt ferrite by varying annealing temperature. Dalt Trans. 2014;43:6919.10.1039/c4dt00166dSearch in Google Scholar PubMed

[147] Praveena K, Srinath S. Synthesis and characterization of CoFe2O4/polyaniline nanocomposites for electromagnetic interference applications. J Nanosci Nanotechnol. 2014;14:4371–6.10.1166/jnn.2014.8286Search in Google Scholar PubMed

[148] Xu ST, Ma YQ, Zheng GH, Dai ZX. Simultaneous effects of surface spins: rarely large coercivity, high remanence magnetization and jumps in the hysteresis loops observed in CoFe2O4 nanoparticles. Nanoscale. 2015;7:6520–6.10.1039/C5NR00582ESearch in Google Scholar PubMed

[149] Sattarahmady N, Zare T, Mehdizadeh AR, Azarpira N, Heidari M, Lotfi M, et al. Dextrin-coated zinc substituted cobalt-ferrite nanoparticles as an MRI contrast agent: in vitro and in vivo imaging studies. Colloids Surf B Biointerf. 2015;129:15–20.10.1016/j.colsurfb.2015.03.021Search in Google Scholar PubMed

[150] Alves TEP, Pessoni HVS, Franco A. The effect of Y3+ substitution on the structural, optical band-gap, and magnetic properties of cobalt ferrite nanoparticles. Phys Chem Chem Phys. 2017;19:16395–405.10.1039/C7CP02167DSearch in Google Scholar PubMed

[151] Mirzaee S, Shayesteh SF. Ultrasound induced strain in ultrasmall CoFe2O4@polyvinyl alcohol nanocomposites. Ultrason Sonochem. 2018;40:583–6.10.1016/j.ultsonch.2017.07.050Search in Google Scholar

[152] Yadav RS, Kuřitka I, Vilcakova J, Havlica J, Kalina L, Urbánek P, et al. Sonochemical synthesis of Gd3+ doped CoFe2O4 spinel ferrite nanoparticles and its physical properties. Ultrason Sonochem. 2018;40:773–83.10.1016/j.ultsonch.2017.08.024Search in Google Scholar

[153] Cushing BL, Kolesnichenko VL, O’Connor C. Recent advances in the liquid-phase syntheses of inorganic nanoparticles. J Chem Rev. 2004;104:3893–946.10.1021/cr030027bSearch in Google Scholar

[154] Jiu J, Ge Y, Li X, Nie L. Preparation of Co3O4 nanoparticles by a polymer combustion route. Mater Lett. 2002;54:260–3.10.1016/S0167-577X(01)00573-0Search in Google Scholar

[155] Shinde AB. Structural and electrical properties of cobalt ferrite nanoparticles. Int J Innov Technol Explor Eng. 2013;3:64.Search in Google Scholar

Received: 2020-11-18
Accepted: 2020-12-09
Published Online: 2020-12-31

© 2020 Claudia Petrarca et al., published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 16.4.2024 from https://www.degruyter.com/document/doi/10.1515/ntrev-2020-0111/html
Scroll to top button