Next Article in Journal
A Study of Physical Modeling and Mathematical Modeling on Inclusion Behavior in a Planar Flow Casting Process
Next Article in Special Issue
Review on Grain Refinement of Metallic Materials to Regulate Cellular Behavior
Previous Article in Journal
In Vitro Corrosion Performance of As-Extruded Mg–Gd–Dy–Zr Alloys for Potential Orthopedic Applications
Previous Article in Special Issue
Influence of Pr3+ and CO32− Ions Coupled Substitution on Structural, Optical and Antibacterial Properties of Fluorapatite Nanopowders Obtained by Precipitation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

An Overview of the Production of Magnetic Core-Shell Nanoparticles and Their Biomedical Applications

by
Dimitris Tsamos
1,*,
Athina Krestou
2,
Maria Papagiannaki
3 and
Stergios Maropoulos
4,*
1
Laboratory of Biomaterials and Computational Mechanics, Department of Mechanical Engineering, University of Western Macedonia, 50100 Kozani, Greece
2
Department of Mechanical Engineering, University of Western Macedonia, 50100 Kozani, Greece
3
Department of Occupational Therapy, Health Science Faculty, University of Western Macedonia, 50200 Ptolemaida, Greece
4
Laboratory of Mechanical Processes and Quality Control, Centre for Testing of Materials and Constructions, Department of Mechanical Engineering, University of Western Macedonia, 50100 Kozani, Greece
*
Authors to whom correspondence should be addressed.
Metals 2022, 12(4), 605; https://doi.org/10.3390/met12040605
Submission received: 27 February 2022 / Revised: 23 March 2022 / Accepted: 27 March 2022 / Published: 31 March 2022
(This article belongs to the Special Issue Nanoparticles for Biomedical and Cosmetic Applications)

Abstract

:
Several developments have recently emerged for core-shell magnetic nanomaterials, indicating that they are suitable materials for biomedical applications. Their usage in hyperthermia and drug delivery applications has escalated since the use of shell materials and has several beneficial effects for the treatment in question. The shell can protect the magnetic core from oxidation and provide biocompatibility for many materials. Yet, the synthesis of the core-shell materials is a multifaceted challenge as it involves several steps and parallel processes. Although reviews on magnetic core-shell nanoparticles exist, there is a lack of literature that compares the size and shape of magnetic core-shell nanomaterials synthesized via various methods. Therefore, this review outlines the primary synthetic routes for magnetic core-shell nanoparticles, along with the recent advances in magnetic core-shell nanomaterials. As core-shell nanoparticles have been proposed among others as therapeutic nanocarriers, their potential applications in hyperthermia drug delivery are discussed.

1. Introduction

Nanomaterials are of great scientific interest as they possess unique properties compared to their bulk counterparts. Sized in a nanodomain with at least one dimension in the range of a few nanometers (1–100 nm), they usually have size- and shape-dependent physical properties. Among the physical properties reliant on size are superparamagnetism (in magnetic particles), surface plasmon resonance (in several metallic particles), and quantum confinement (in semiconductor particles). While nanomaterials are usually perceived as late inceptions, this is not the case for most of them. The use of nanomaterials dates back to the fourth century AD when Romans started using metallic particles in glass constructions such as cups. The Lycurgus cup is the oldest known case [1], produced from silver and gold nanoparticles (NPs) embedded in glass, just like stained-glass windows. The production process of such materials back then was based on empirical methods as there was a lack of fundamental scientific knowledge. Some of those methods were discovered by an accidental mixture of the glass with pulverized metallic dust. Nowadays, with the evolution of nanotechnology, there is a defined scientific methodology underlining the production of nanoparticles. Additionally, while the production of nanoparticles and thin films (ranging up to 100 nm) has been stressed repeatedly in the literature, there is still a need for updated inclusive reviews, such as the one presented here, focusing on the production of core-shell and nanohybrid particles for biomedical applications.
Furthermore, there has been a staggering increase in the development of complex materials based on various metallic core-shell or nanohybrid particles. Some of them are prominent for their magnetic and electrochemical properties, which render them ideal for applications such as hyperthermia [2], lithium-ion batteries [3], microwave absorption [4], catalysis [5], et cetera. At the same time, other NPs are attractive for their antibacterial and UV-protection properties, which make them ideal for biomedical [6] and cosmetic applications [7,8], respectively. The above stems from their superior biocompatibility as their shell material can be modulated to provide preferable interactions with biological tissue. Their magnetic core material, in return, allows the precise administration of these nanocarriers to a specific anatomical site. The combination of nanocore with nanocoatings renders these particles as advanced materials as it allows application-specific surface functionalization [9], while shielding the core from the particles’ environment. As such, core-shell nanoparticles and nanohybrids are usually multifunctional as their properties can be easily modulated because both the core and the shell of the particle can contribute to their physicochemical characteristics. For instance, their reactivity and thermal stability depends on the coating material and can thus be tailored to meet the application’s requirements. Therefore, the use of coatings on the core materials is manifold, with increased dispersibility, functionality, stability, and surface modification being only some of the elements that can be affected through the applied coating.

2. Scope of this Review

Reviews have an essential role in keeping researchers up to date on the current state of developments in a fairly focused scientific field. While many researchers have also published reviews [10,11,12,13] and book chapters [14,15] on core-shell nanoparticles, they emphasize primarily the applications, the physical and chemical properties, or the specific groups of magnetic materials. Nevertheless, there is a lack of literature that compares the size and shape of magnetic core-shell nanomaterials synthesized via different routes. To this end, the current review provides a systematic overview of the available production methods of magnetic core-shell nanoparticles. As the nanoparticle’s shape and size depend on the production method, the primary aim is to compare the different techniques with respect to the characteristics of the NPs. Consequently, this review can serve as an introduction to core-shell nanoparticle synthesis or as a starting point for researchers seeking alternative production routes for magnetic core-shell nanoparticles, while portraying the importance and future potential of these materials in biomedical and cosmetic applications.

3. Basic Strategies for Producing Nanostructured Materials

Two basic strategies/approaches are used to produce nanoparticles, core-shell nanoparticles, and nanohybrid materials alike: “top-down” and “bottom-up”(Figure 1), with the combination of both strategies being a prevalent route as well. In the top-down approach, a macroscopic material or group of materials decreases from the macroscale to the microscale and then to the nanoscale. With this approach, traditional industrial methods such as microfabrication take place, which use various tools to shape, cut, or resize a bulk material into particles. Several techniques fall within the above approach, with the most prevalent being associated with nano-lithographic techniques such as optical lithography, electron beam lithography, optical near-field lithography, scanning probe lithography, nanoimprint lithography, and others. Besides lithographic techniques, laser-beam processing and several mechanical techniques, such as grinding, milling, machining, and polishing, also produce nanoparticles.
Bottom-up methods utilize the chemical and physical properties of nano-sized molecules to drive their self-assembly and to grow functional structures. This route produces distinct, manifold, and complex structures from atoms or molecules, at high precision, and results in refined control over the shape and size of the NPs. Over the past decades, many bottom-up strategies have been introduced, extending from condensing vapors on surfaces to the coalescence of atoms in liquid solutions.
Some of the most prevalent bottom-up methods are aerosol processes, chemical synthesis (including precipitation techniques, reduction reactions, sol-gel processes, hydrothermal and solvothermal reactions, et cetera), molecular self-assembly, laser-induced assembly, chemical and physical vapor deposition techniques, et cetera.

4. Top-Down Processes

4.1. Milling Processes

Those processes produce nanoparticles by sequentially reducing the size of bulk materials into smaller, yet stable structures. The above is achievable by the interacting of the raw materials (usually a powder with a grain size of several μm) with a mechanical device called a mill. These powders undergo plastic deformation through a series of impacts that lead to their fracture, ultimately reducing their particle size. Consequently, the size can decrease even below 100 nm, producing nanoparticles. The milling processes produce nanocrystalline materials in solid-state at standard ambient conditions, using inexpensive equipment, and they are fully scalable to the industrial level. However, the disadvantage of the milling process is material contamination as lengthy processing and high temperatures result in the diffusion of materials from the grinding medium and the container surface. In addition, milling does not comprehensively control the particle shape and yields powders with relatively broad particle-size ranges. Some of the above drawbacks can be eliminated through wet milling, where a small amount of organic solvent and surfactants is added to the mixture to stabilize the nanoparticles formed by the impact and prevent their agglomeration.
High-energy milling techniques, such as vibrating, planetary, and friction milling, are usually preferred, with planetary high-energy ball milling being the most prominent. High-energy ball milling contains a low number of balls (Figure 2), functioning as the grinding medium inside the container, that move freely and have adequate space to gain a high momentum before any impact. A motor that shakes the container in a planetary motion provides the system with high energy levels capable of milling even the most challenging materials. Usually, those mills have both a container and grinding balls composed of steel or tungsten carbide.

4.2. Lithographic Processes

In material science, the term lithographic processes refers to the manufacturing techniques for inducing desired patterns in micro- and nanoscale. Therefore, they are usually referred to as microlithography or nanolithography, depending on the target size scale of the process. While the above processes are typical for producing nanostructured patterns, they can also produce nanoparticles [16]. In lithographic techniques, a distinct procedure is used to mark thin films of barrier materials that cover substrates during the deposition, implantation, or etching procedures. In contrast with other methods, lithography provides consistency and full control of the pattern’s shape and size but requires costly equipment and demanding infrastructure. Well-defined nanoparticles can be fabricated by lithographic methods independently or with the assistance of other processes, such as physical vapor deposition [17] (PVD), reactive ion etching [18], and other assorted methods. Some of the most popular lithographic methods used to produce nanostructures and nanoparticles are photolithography, electron beam lithography, nanoimprint lithography, interference lithography, nanosphere lithography, nanostencil lithography, and others.
The semiconductor industry widely uses lithographic processes, especially photolithography, to produce state-of-the-art semiconductors and, consequently, electronics [19]. For the above reason, photolithography is one of the most mature techniques, maintaining high reliability. Photolithography or optical lithography uses light (usually UV radiation) and a mask to sculpture a geometric pattern on a light-sensitive coating of a substrate. Depending on its chemistry, the light-sensitive material disintegrates or hardens due to exposure, while solvents remove the excess coating, creating a pattern. The remaining pattern layer acts as a guide to producing the desired micro- and nanostructures via deposition, etching, and other procedures. The radiation’s wavelength defines the minimum feature size to engrave the photoresist layer.
Interference or holographic lithography is a photolithography variation that does not require the usage of masks. Instead, it uses short exposure of coherent light waves to produce nanoscale patterns in a photoresist over a large area. Two coherent beams, created from a laser source by a beam splitter, are piloted to superimpose and interfere with one another, creating a spatial interference pattern consisting of periodically distributed intensity on the photoresist.
Apart from interference lithography, electron beam lithography is a common mask-less technique. In the above process, a focused beam of electrons is used to mark a layer of an electron-sensitive film, anointed as resist, changing its solubility. Then, the resist’s exposed or non-exposed areas can be removed using a solvent. Eventually, evaporation of the desired material on the patterned resist can produce functional nanoparticles, and a lift-off process can clear the resist. It is also attainable by transferring the patterns to the functional layer with dry or wet etching processes.
The Figure 3 below illustrates photolithography and electron beam lithography.

5. Bottom-Up/Chemical Production Processes

5.1. Sol-Gel

The colloidal gel technique (sol-gel) is a wet chemical process initially used to synthesize inorganic materials such as glass, ceramics, ceramic nanostructured polymers, porous nanomaterials, and oxide nanoparticles. The low temperature and mild conditions required for this method enable the synthesis of hybrid coatings from organic/inorganic materials incorporating low molecular weight organic compounds. The sol-gel process mainly transforms a colloidal liquid state of dispersed particles (sol) to a solid-state (gel) interconnected network of polymer chains with an average length of more than one micrometer and a pore size of less than one micrometer. The precursors used for sol-gel synthesis are metals or metalloids attached with organic compounds. Metal alkoxides (metallic compounds of alcohol, with silicon, titanium, or aluminum) are the most prevalent because of their high reactivity with water.
Inserting organic substances into the wetting process produces an organometallic compound from an alkoxide solution. Subsequently, the solution’s pH is adjusted with an acidic or basic solution. Next to the pH regulation, these media also work as catalysts, triggering the transformation of the alkoxide. The reactions involved in this are hydrolysis, followed by condensation and polymerization. The polymer oxide or the particles grow as the reaction proceeds until a gel is formed. It is noteworthy that the hydrolysis and the condensation reaction depend on the initial solution’s composition, catalyst volume, temperature, and reactor-mixing geometry. In the case of coatings, the alkoxide solution is applied even to geometrically complex surfaces. Once applied to a surface, the growth of a porous network materializes via the gel formation. Thicker layers can be produced by repeated wetting and drying. A drying step always follows the gel formations (wetting).
With the ability to process every step separately in the formation of the sol and gels, it is possible to modify the entire technique to produce powders, fibers, ceramics, coatings, and even highly porous nanomaterials (Figure 4). Furthermore, the porous network created from the gel step allows the production of new composites by filling the pores during or after the gelling process, due to the low operating temperature. However, one major drawback of the sol-gel processes is the controlling of the synthesis and drying steps, combined with the organic contaminants that remain in the gel. Ultimately, thermal post-treatment is usually required, adding to the complexity of scaling up those processes.
Over the years, multiple sol-gel variations have emerged, with the Stöber process being one of the most efficient [20]. The Stöber method is a sol-gel process in which the molecular precursor (usually tetraethylorthosilicate) reacts first with water in an alcohol solution; then, the consequent molecules agglomerate to build larger structures. The aforementioned method is employed to prepare SiO2 nanoparticles in a controllable shape and size, and it is quite an efficient sol-gel variation.

5.2. Co-Precipitation

This technique is relatively straightforward for nanoparticle synthesis and involves dissolving metal precursors in a polar solvent, usually water, along with a reducing agent. AS a polar solvent is utilized, the most common precursors are ionic compounds such as metal salts. The process is quite efficient and lasts from a few minutes to a few hours. The cores of the nanoparticles are formed from the reduction of the metals. Thereupon, the particles settle as a precipitate in the solution. At the final stage of the procedure, the particles are collected by ultra-centrifugation or by a magnet in the case of magnetic materials and thermally post-treated. The nanoparticles’ size, shape, crystallinity, and composition depend on the type of salts (chlorides, sulfates, and nitrates), the ionic compound ratio, the reaction temperature, the pH, and the ionic strength of the solvent [21]. This method has the edge on high repeatability and the quantities of nanoparticles produced. However, it is challenging to control the shape in this approach because only the kinetic factors are modified during synthesis. Furthermore, higher temperatures would lead to agglomeration succeeding in destroying the nanostructure. Therefore, co-precipitation remains one of the most frequently utilized production processes for nanomaterials.

5.3. Emulsions

Those processes involve a thermodynamically stable dispersion of two immiscible liquids, usually oil and water, with the presence of a surfactant. The surfactant stabilizes the interface between the two liquid phases. For example, in oil and water microemulsions, the organic phase is dispersed in microdroplets, enclosed by a layer of surfactant molecules, separating them from the aqueous phase. This separation occurs so that the surfactant’s hydrophilic part is inside the aqueous environment, while the hydrophobic part is inside the organic phase, resulting in the formation of micelles. The molecular fraction of the aqueous phase to the surfactant and the surfactant’s chemical structure defines the micelle’s size and structure. Fundamentally, micelles are formed due to amphiphilic (containing both hydrophobic and hydrophilic groups) surfactants, as shown in Figure 5. The surfactants, and the micelles that they create, are excellent tools of reaction that can work as nanoreactors as they manage to increase the solubility of the reactants and organize them within their structure. The above is achievable as micelles deliver constant mobility when they collide, unite, and redissolve. Consequently, a precipitate eventually forms inside them and is finally isolated by centrifugation, filtration, or magnetic separation in post-treatment. There are two types of micelles, regular and reverse micelles. In the latter, the aqueous phase is trapped in droplets inside the organic phase.
Even though emulsions can synthesize multiple types of nanoparticles, they do not deliver reasonable control over size and shape while maintaining small production yields. Nevertheless, it is worth noting that while emulsion processes can produce magnetic nanoparticles for various applications [22], they are among the preferred production routes for nanoparticles in biological [23,24] and cosmetic applications [25].

5.4. Hydrothermal—Solvothermal

Hydrothermal synthesis refers to the heterogeneous reaction in an aqueous solution inside a closed system (autoclave) when the temperature and the pressure exceed 100 °C and 1 bar, respectively. The autoclave (a high-pressure steel container, as shown in Figure 6) essentially ensures controlled crystal growth. The above is due to the high pressure favoring the formation of complex structures with the desired crystallinity at a relatively low temperature. In addition, one quite distinctive advantage is that hydrothermal synthesis is a straightforward process with low cost, environmentally friendly reagents, and satisfactory repeatability, despite its low production yield. If an organic solvent is used instead of water, the process is termed solvothermal. A critical difference between the two methods is the pressure inside the autoclave. In hydrothermal synthesis, the system’s pressure is determined externally, while in solvothermal it is autogenous (self-dependent) and cannot be modified.
Furthermore, it is typical for the hydrothermal and solvothermal synthesis of inorganic nanoparticles to occur at higher temperatures than the solvent’s boiling point. In this way, the solvent is in a supercritical state, and its properties (density, surface tension, viscosity, and dielectric constant) change drastically, greatly facilitating the consummation of the reactions. Moreover, the solvent’s properties vary between gases and liquids in their supercritical state. Therefore, even though accurate knowledge of the reactions during hydrothermal and solvothermal synthesis is complicated, they can be classified merely into redox reactions, hydrolysis, thermolysis, complex formation reactions, and permutation reactions. The parameters that affect the synthesis fall into two categories: chemical and thermodynamic. The chemical parameters depend on the nature of the reactants and the solvent. In addition, the chemical composition and concentration of the precursors in the solution affect the characteristics of the nanoparticles produced. Thus, choosing the correct solvent takes effect by assessing the molecular weight, boiling point, dielectric constant, bipolar moment, and polarity of the solvent. The thermodynamic parameters include the temperature, the pressure, and the reaction time, which affect the morphology and the composition of the NPs. Eventually, any mixture of reactants and solvents under precise experimental conditions can lead to new nanoparticles.

5.5. Thermal Decomposition

Thermal decomposition is the method that utilizes the high-temperature decomposition of precursors, such as salts, metal, and organometallic compounds, through a series of chemical reactions. These reactions occur in organic solvents in the presence of surfactants, which are often fatty acids such as oleic acid or amines. As the process mandates an inert atmosphere, the equipment used is more complex and therefore more expensive than the equipment used on other techniques, such as co-precipitation or sol-gel. Furthermore, the produced nanoparticles are mainly hydrophobic; so, their surface needs further processing for biomedical applications. The morphology of the produced NPs depends on the concentrations of the precursors and the solvent, temperature, and reaction time, whereas the pressure is autogenous and, therefore, is not controllable. The main advantage of thermal decomposition over other bottom-up methodologies is the narrow size distribution and the shape homogeneity of the produced nanoparticles. However, one major drawback of the process is the poor repeatability due to numerous parameters affecting the synthesis.
Thermal decomposition is usually based on a hot-injection technique, where the precursor compounds are added to a pre-heated solvent and surfactant solution [21]. The above leads to an abrupt nucleation, separating the growth stage from the nucleation. The apparatus utilized in this procedure is illustrated in the figure below, revealing that it is possible to insert components during the reaction through the right nozzle of the three-nozzle spherical flask.
Figure 7 portrays the hot-injection setup that allows the control of the reactions to a certain degree, while also facilitating visual observations, such as the solution’s color change. In addition, the ability to insert precursors during the proceeding reaction allows the separation of the nucleation into two steps, with the second step initiating after the intake of new precursors. In this manner, it is possible to prepare heterometallic nanoparticles, with the second metal compound infused at a specific reaction stage.

5.6. Gas-Phase—Vacuum Processes

Gas-phase methods are industry-standard techniques for producing nanomaterials in powder or film form. However, most of them require an inert atmosphere or even vacuum pressures to deliver the desired results, which significantly increases the cost of equipment and consumables. Nevertheless, the production consistency and high purity of the synthesized NPs have rendered these techniques highly popular. In gas-phase processes, the vapor from the precursor compounds produces nanomaterials employing either chemical or physical mechanisms. Nanoparticles are mainly synthesized in liquid or solid state through homogeneous nucleation, with various additional mechanisms such as chemical reactions on the surface, adhesion of two or more particles, particle fusion, and condensation leading to the final nanomaterials. Some of the most frequent gas-phase techniques involve hot-wall, plasma, laser, and flame reactors, with Chemical Vapor Deposition (CVD) being the most widely used of all of them.
Chemical Vapor Deposition is a vacuum deposition technique that produces materials through the chemical reactions of gaseous reagents. Usually, the substrate is exposed to gaseous precursors and solid materials (thin films or nanoparticles) produced along with volatile by-products, which are removed from the reactor with pumps. Despite several variations, all CVD processes require an energy source to break down the gaseous reagents and initiate the chemical reaction. The substrate can be heated with either one of two methods: directly or indirectly (Figure 8). In direct methods, the substrate absorbs energy either inductively or through another heating source at its base surface. This type of CVD employs cold-wall reactors, and the chamber wall remains at ambient temperature. In indirect methods, the entire chamber is externally heated, with the substrate temperature increasing through radiation from the chamber, which is part of a hot-wall reactor. CVD processes usually involve high temperatures and low pressures. Depending on the pressure deployed in the reaction chamber, CVD is classified into the following types: Ultra-High Vacuum CVD (UHVCVD), where the pressure drops below 10−8 Torr and Low-Pressure CVD (LPCVD), with the chamber pressure ranging between 10−2 -10 Torr. Notably, in low-pressure conditions, the unwanted gas-phase reactions are restricted and the film uniformity across the substrate is improved, whereas UHVCVD creates high-purity products. Another less popular type of CVD is atmospheric pressure CVD.
CVD processes have numerous variations, such as the plasma-enhanced CVD, laser CVD, metal-organic CVD, and others that fall beyond the scope of this review. However, chemical vapor deposition techniques deliver high-quality, high-performance, and high-purity solid nanomaterials. As a result, they are popular in the semiconductor industry, where they are employed to produce semiconductors, metals, oxides, nitrides, and organic nanomaterials.

6. Concentrated Table

The following table includes the most popular synthesis methods for magnetic core-shell nanoparticles. It is noteworthy that magnetic core-shell nanoparticles are usually multifunctional. Even though not all the articles cited in Table 1 refer to core-shell particles utilized for biomedical applications, even the methodologies presented for the non-biomedical ones can be extended to purposes that fall within the scope of this review if combined with a proper shell. For instance, iron oxide nanoparticles can be utilized for lithium-ion batteries [3], microwave absorption [4], catalysis [5], and even for wastewater treatment [9] and many more applications, with the synthesis method remaining unaltered across different applications. Therefore, the following table contains core-shell nanoparticles for various applications. Its main objective is to provide a handy tool for comparing different sizes, shapes, and synthetic methods on various magnetic nanoparticles.

7. Biomedical Applications

Nanomaterials have been suggested for a variety of bio-related applications, whether for medical applications [95,96], agro-biotechnology [97], or cosmetic ones [98]. Regulatory concerns have, however, so far restricted their wide use to high added-value applications, conventionally found in biomedical applications. Among these are therapeutic delivery systems, to which nanomaterials have been suggested since the early 90s [99,100], with core-shell nanoparticles following only a few years after [101]. However, as far as it concerns biomedical or pharmaceutical applications, the main advantage of core-shell nanoparticles over other materials ranging in the nano-domain, is affiliated with their biocompatibility [102]. The above can be attributed to the combination of a magnetic core with bio-inspired shells, ensuring favorable interaction with biological systems.
The latest research regarding these nanocarriers has focused mainly on diseases where conventional medication yields a low therapeutic index, i.e., cancer treatment [103]. Next to targeted drug delivery, core-shell nanoparticles have elicited attention due to their superparamagnetic properties, which they exhibit in the presence of a magnetic core (e.g., iron oxides and sulfides). The advantage offered by ferrous core material is bifold: (a) administering the nanocarrier to a specific anatomical site (or even target cell) through the application of an external magnetic field that navigates the particles as desired [104] and (b) through their exposure to an oscillating magnetic field, which increases their temperature and induces controlled hyperthermia [105]. Hyperthermia has lately emerged as a valid alternative to radiotherapy and chemotherapy, as malignant cells are prone to temperatures beyond 41 °C. Furthermore, the site-specific heat administration, along with the restriction of temperature ranges to below 44 °C, can preserve the surrounding tissue far more effectively than other oncotherapies [106].

7.1. Hyperthermia

While hyperthermia is a rather old concept, introduced with the injecting of iron oxide nanoparticles into lymphatic channels [107], nowadays it has evolved rapidly. Modern research has progressed significantly in improving efficiency and productions methods.
Interestingly, Simeonidis et al. [33] recently reported a heating efficiency of approximately 0.9 kW/g in large spherical iron/iron-oxides core−shell nanoparticles. This was achieved by exploiting the interactions between the core and shell phases. Their group proved that the final particle size, the core−shell ratio, and the interposition of a thin wüstite interlayer are correlated to specific absorption rates. They used Solar Vapor Phase Condensation (Solar PVD) as a green, cost-effective, and scalable technique, that can produce large quantities of dry nanoparticle powders with a narrow size distribution without the purification steps.
Apart from the previously mentioned progress, recent advances in the production of magnetic materials for hyperthermia include improvements in the synthesis of the popular silica-coated nanoscale zero-valent iron oxides. Specifically, Tajabadi et al. [108] managed to optimize the synthesis, revising the molar ratio of the precursors; using response surface methodology, they created various modified samples with a high maximum saturation magnetization of 99.3 emu/g. Furthermore, they achieved SAR values of 354.7 and 419.9 W/g for the most optimized samples, which are relatively high for Fe@SiO2 nanoparticles. In addition, the above samples had enough silica coating to protect the inner core from erosion and prevent the release of iron ions. Yet, the core-shell particles with a dense silica shell revealed an acceptable cytocompatibility of up to 250 lg/mL. It is noteworthy that all the above optimizations were possible with synthesis simulation techniques such as response surface methodology.

7.2. Drug Delivery

Several novel concepts other than hyperthermia have come to light as numerous research groups globally investigate drug delivery. Some include magnetic core/shell nanoparticles that combine photo-induced hyperthermia and drug delivery [65]. For example, in the above study, Adam et al. produced iron oxide@stellate mesoporous silica nanocarriers to deliver doxorubicin (a chemotherapy medication used to treat cancer). The photothermia was induced through near-infrared light and correlated to the energy of the applied field. Moreover, the pH-induced delivery of the antitumor drug was then tailored through the surface functionalization of their core-shell systems. In particular, the mesoporous silica’s stellate shape (Figure 9), the surface functionalization, and the pH tuning made the doxorubicin loading possible to an extended degree. Therefore, with the nanoparticles mentioned above, it is possible to include drug delivery and hyperthermia applications simultaneously.
In a similar study, Chen et al. [109] utilized a microfluidic technology to synthesize composite materials composed of a thermosensitive hydrogel with incorporated magnetic core/shell nanoparticles. These crosslinked polymers were used as single emulsions, encapsulating an aqueous drug or double emulsions loaded with a water-soluble drug in the shell and an oil-soluble one in the core. The core and shell layers of the emulsions are filled with two drugs, camptothecin and doxorubicin hydrochloride, to perform simultaneous delivery of hydrophobic and hydrophilic medicines. In addition, the drug delivery of the composites is possible by employing Fe3O4 nanoparticles in the shell layer of the emulsions. Finally, the magnetic field-induced hyperthermia was used, along with its anti-cancer properties, to increase the temperature of the hydrogel in a controlled manner, thus dictating a targeted medicinal effect through the controlled release of the encapsulated drugs.

8. Concluding Remarks and Perspective

Multifunctional magnetic core-shell nanoparticles are emerging platforms for many applications. However, despite a plethora of available fabrication routes that can be employed for their production, some (e.g., co-precipitation or reduction of the core) are favored more than others as they are more productive and cost-efficient.
The acceptance of core-shell nanoparticles from the biomedical sector was driven mainly by their exceptional (shell-based) biocompatibility, combined with their outstanding guidance potential and traceability, based on their magnetic cores. In addition, while some biomedical applications in drug delivery [101] and hyperthermia [107] are decade-old concepts, their research output has escalated to promising clinical trials [110]. The above is valid due to the favorable biocompatibility of core-shell nanoparticles when compared to other nanomaterials [102], usually attributed to the composition of their shell. Despite the numerous challenges, magnetic core-shell nanoparticles are regarded as a promising class of materials for many other research areas, e.g., catalysis, radio absorption, and energy storage systems. Therefore, efficient and controlled production methods for core-shell nanoparticles continue to pose a scientific challenge.

Author Contributions

Conceptualization, D.T., M.P., A.K. and S.M.; data analysis, D.T. and M.P.; writing—original draft preparation, D.T. and M.P.; review and editing, A.K. and S.M.; supervision, A.K. and S.M.; funding acquisition, S.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research has been implemented in the framework of the Action “Support for Research, Technological Development and Innovation Projects in RIS3 Sectors” and co-financed by the European Union and national resources through the CENTRAL GREECE 2014-2020 (project code: SΤΕΡ1-0025339).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Freestone, I.; Meeks, N.; Sax, M.; Higgitt, C. The Lycurgus Cup—A Roman Nanotechnology. Gold Bull. 2007, 40, 270–277. [Google Scholar] [CrossRef] [Green Version]
  2. Christodoulou, E.; Nerantzaki, M.; Nanaki, S.; Barmpalexis, P.; Giannousi, K.; Dendrinou-Samara, C.; Angelakeris, M.; Gounari, E.; Anastasiou, A.D.; Bikiaris, D.N. Paclitaxel Magnetic Core–Shell Nanoparticles Based on Poly(Lactic Acid) Semitelechelic Novel Block Copolymers for Combined Hyperthermia and Chemotherapy Treatment of Cancer. Pharmaceutics 2019, 11, 213. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Kang, E.; Jung, Y.S.; Cavanagh, A.S.; Kim, G.-H.; George, S.M.; Dillon, A.C.; Kim, J.K.; Lee, J. Fe3O4 Nanoparticles Confined in Mesocellular Carbon Foam for High Performance Anode Materials for Lithium-Ion Batteries. Adv. Funct. Mater. 2011, 21, 2430–2438. [Google Scholar] [CrossRef]
  4. Wang, T.; Liu, Z.; Lu, M.; Wen, B.; Ouyang, Q.; Chen, Y.; Zhu, C.; Gao, P.; Li, C.; Cao, M.; et al. Graphene–Fe3O4 Nanohybrids: Synthesis and Excellent Electromagnetic Absorption Properties. J. Appl. Phys. 2013, 113, 024314. [Google Scholar] [CrossRef]
  5. Wu, Z.-S.; Yang, S.; Sun, Y.; Parvez, K.; Feng, X.; Müllen, K. 3D Nitrogen-Doped Graphene Aerogel-Supported Fe3O4 Nanoparticles as Efficient Electrocatalysts for the Oxygen Reduction Reaction. J. Am. Chem. Soc. 2012, 134, 9082–9085. [Google Scholar] [CrossRef] [PubMed]
  6. Kirmanidou, Y.; Sidira, M.; Bakopoulou, A.; Tsouknidas, A.; Prymak, O.; Papi, R.; Choli-Papadopoulou, T.; Epple, M.; Michailidis, N.; Koidis, P.; et al. Assessment of Cytotoxicity and Antibacterial Effects of Silver Nanoparticle-Doped Titanium Alloy Surfaces. Dent. Mater. 2019, 35, e220–e233. [Google Scholar] [CrossRef] [PubMed]
  7. Regulation (EC) No 1223/2009 of 30 November 2009 on Cosmetic Products. Off. J. Eur. Union 2009, 342, 59–209.
  8. Karamanidou, T.; Bourganis, V.; Gatzogianni, G.; Tsouknidas, A. A Review of the EU’s Regulatory Framework for the Production of Nano-Enhanced Cosmetics. Metals 2021, 11, 455. [Google Scholar] [CrossRef]
  9. Radoń, A.; Łoński, S.; Kądziołka-Gaweł, M.; Gębara, P.; Lis, M.; Łukowiec, D.; Babilas, R. Influence of Magnetite Nanoparticles Surface Dissolution, Stabilization and Functionalization by Malonic Acid on the Catalytic Activity, Magnetic and Electrical Properties. Colloids Surf. A Physicochem. Eng. Asp. 2020, 607, 125446. [Google Scholar] [CrossRef]
  10. Nnadozie, E.C.; Ajibade, P.A. Multifunctional Magnetic Oxide Nanoparticle (MNP) Core-Shell: Review of Synthesis, Structural Studies and Application for Wastewater Treatment. Molecules 2020, 25, 4110. [Google Scholar] [CrossRef] [PubMed]
  11. Katz, E. Synthesis, Properties and Applications of Magnetic Nanoparticles and Nanowires—A Brief Introduction. Magnetochemistry 2019, 5, 61. [Google Scholar] [CrossRef] [Green Version]
  12. Kwizera, E.A.; Chaffin, E.; Wang, Y.; Huang, X. Synthesis and Properties of Magnetic-Optical Core–Shell Nanoparticles. RSC Adv. 2017, 7, 17137–17153. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Hao, R.; Xing, R.; Xu, Z.; Hou, Y.; Gao, S.; Sun, S. Synthesis, Functionalization, and Biomedical Applications of Multifunctional Magnetic Nanoparticles. Adv. Mater. 2010, 22, 2729–2742. [Google Scholar] [CrossRef] [PubMed]
  14. El-Gendy, A.A. Core/Shell Magnetic Nanoparticles for Biomedical Applications. In Magnetic Nanostructured Materials; Elsevier: Amsterdam, The Netherlands, 2018; pp. 41–58. [Google Scholar]
  15. Mandal, S.; Chaudhuri, K. Magnetic Core-Shell Nanoparticles for Biomedical Applications. In Complex Magnetic Nanostructures; Springer International Publishing: Cham, Switzerland, 2017; pp. 425–453. [Google Scholar]
  16. Fu, X.; Cai, J.; Zhang, X.; Li, W.-D.; Ge, H.; Hu, Y. Top-down Fabrication of Shape-Controlled, Monodisperse Nanoparticles for Biomedical Applications. Adv. Drug Deliv. Rev. 2018, 132, 169–187. [Google Scholar] [CrossRef] [PubMed]
  17. Handbook of Physical Vapor Deposition (PVD) Processing; Elsevier: Amsterdam, The Netherlands, 2010; ISBN 9780815520375.
  18. Tachi, S.; Tsujimoto, K.; Okudaira, S. Low-temperature Reactive Ion Etching and Microwave Plasma Etching of Silicon. Appl. Phys. Lett. 1988, 52, 616–618. [Google Scholar] [CrossRef]
  19. Van Zant, P. Microchip Fabrication, Sixth Edition: A Practical Guide to Semiconductor Processing, 6th ed.; McGraw-Hill Education: New York, NY, USA, 2013. [Google Scholar]
  20. Stöber, W.; Fink, A.; Bohn, E. Controlled Growth of Monodisperse Silica Spheres in the Micron Size Range. J. Colloid Interface Sci. 1968, 26, 62–69. [Google Scholar] [CrossRef]
  21. Radoń, A.; Kądziołka-Gaweł, M.; Łukowiec, D.; Gębara, P.; Cesarz-Andraczke, K.; Kolano-Burian, A.; Włodarczyk, P.; Polak, M.; Babilas, R. Influence of Magnetite Nanoparticles Shape and Spontaneous Surface Oxidation on the Electron Transport Mechanism. Materials 2021, 14, 5241. [Google Scholar] [CrossRef]
  22. Lu, A.-H.; Salabas, E.L.; Schüth, F. Magnetic Nanoparticles: Synthesis, Protection, Functionalization, and Application. Angew. Chem. Int. Ed. 2007, 46, 1222–1244. [Google Scholar] [CrossRef]
  23. Jampílek, J.; Kráľová, K.; Campos, E.V.R.; Fraceto, L.F. Bio-Based Nanoemulsion Formulations Applicable in Agriculture, Medicine, and Food Industry. In Nanobiotechnology in Bioformulations; Springer: Berlin/Heidelberg, Germany, 2019; pp. 33–84. [Google Scholar]
  24. Pavoni, L.; Perinelli, D.R.; Bonacucina, G.; Cespi, M.; Palmieri, G.F. An Overview of Micro- and Nanoemulsions as Vehicles for Essential Oils: Formulation, Preparation and Stability. Nanomaterials 2020, 10, 135. [Google Scholar] [CrossRef] [Green Version]
  25. Che Marzuki, N.H.; Wahab, R.A.; Abdul Hamid, M. An Overview of Nanoemulsion: Concepts of Development and Cosmeceutical Applications. Biotechnol. Biotechnol. Equip. 2019, 33, 779–797. [Google Scholar] [CrossRef] [Green Version]
  26. Cervera, L.; Peréz-Landazábal, J.I.; Garaio, E.; Monteserín, M.; Larumbe, S.; Martín, F.; Gómez-Polo, C. Fe-C Nanoparticles Obtained from Thermal Decomposition Employing Sugars as Reducing Agents. J. Alloys Compd. 2021, 863, 158065. [Google Scholar] [CrossRef]
  27. Shinde, K.P.; Ranot, M.; Choi, C.J.; Kim, H.S.; Chung, K.C. Plasma-Assisted Synthesis and Study of Structural and Magnetic Properties of Fe/C Core Shell. AIP Adv. 2017, 7, 075013. [Google Scholar] [CrossRef]
  28. Cao, H.; Huang, G.; Xuan, S.; Wu, Q.; Gu, F.; Li, C. Synthesis and Characterization of Carbon-Coated Iron Core/Shell Nanostructures. J. Alloys Compd. 2008, 448, 272–276. [Google Scholar] [CrossRef]
  29. Ma, C.; Luo, B.; Song, H.; Zhi, L. Preparation of Carbon-Encapsulated Metal Magnetic Nanoparticles by an Instant Pyrolysis Method. New Carbon Mater. 2010, 25, 199–204. [Google Scholar] [CrossRef]
  30. Byeon, J.H.; Kim, Y.-W. Hybrid Gas-Phase Synthesis of Nanoscale Fe–SiO2 Core–Shell Agglomerates for Efficient Transfection into Cells and Use in Magnetic Cell Patterning. RSC Adv. 2013, 3, 13685–13689. [Google Scholar] [CrossRef]
  31. Yuan, M.; Tao, J.; Yan, G.; Tan, M.; Qiu, G. Preparation and Characterization of Fe/SiO2 Core/Shell Nanocomposites. Trans. Nonferrous Met. Soc. China 2010, 20, 632–636. [Google Scholar] [CrossRef]
  32. Zhang, X.F.; Dong, X.L.; Huang, H.; Lv, B.; Zhu, X.G.; Lei, J.P.; Ma, S.; Liu, W.; Zhang, Z.D. Synthesis, Structure and Magnetic Properties of SiO2-Coated Fe Nanocapsules. Mater. Sci. Eng. A 2007, 454–455, 211–215. [Google Scholar] [CrossRef]
  33. Simeonidis, K.; Martinez-Boubeta, C.; Serantes, D.; Ruta, S.; Chubykalo-Fesenko, O.; Chantrell, R.; Oró-Solé, J.; LI. Balcells, A.; Kamzin, A.S.; Nazipov, R.A.; et al. Controlling Magnetization Reversal and Hyperthermia Efficiency in Core–Shell Iron–Iron Oxide Magnetic Nanoparticles by Tuning the Interphase Coupling. ACS Appl. Nano Mater. 2020, 3, 4465–4476. [Google Scholar] [CrossRef]
  34. Durgesh; Sharma, A.; Aggarwal, S.; Kumar, R. Effect of Incorporation of Synthesized Fe @ Ag Core-Shell Nanoparticles on Optical Parameters of Polyvinyl Alcohol. API Conf. Proc. 2019, 2093, 020017. [Google Scholar]
  35. Lu, L.; Zhang, W.; Wang, D.; Xu, X.; Miao, J.; Jiang, Y. Fe@Ag Core–Shell Nanoparticles with Both Sensitive Plasmonic Properties and Tunable Magnetism. Mater. Lett. 2010, 64, 1732–1734. [Google Scholar] [CrossRef]
  36. Malik, M.A.; Alshehri, A.A.; Patel, R. Facile One-Pot Green Synthesis of Ag–Fe Bimetallic Nanoparticles and Their Catalytic Capability for 4-Nitrophenol Reduction. J. Mater. Res. Technol. 2021, 12, 455–470. [Google Scholar] [CrossRef]
  37. Zhang, X.-B.; Yan, J.-M.; Han, S.; Shioyama, H.; Xu, Q. Magnetically Recyclable Fe@Pt Core−Shell Nanoparticles and Their Use as Electrocatalysts for Ammonia Borane Oxidation: The Role of Crystallinity of the Core. J. Am. Chem. Soc. 2009, 131, 2778–2779. [Google Scholar] [CrossRef] [PubMed]
  38. Pisane, K.L.; Singh, S.; Seehra, M.S. Synthesis, Structural Characterization and Magnetic Properties of Fe/Pt Core-Shell Nanoparticles. J. Appl. Phys. 2015, 117, 17D708. [Google Scholar] [CrossRef] [Green Version]
  39. Burke, N.A.D.; Stöver, H.D.H.; Dawson, F.P. Magnetic Nanocomposites: Preparation and Characterization of Polymer-Coated Iron Nanoparticles. Chem. Mater. 2002, 14, 4752–4761. [Google Scholar] [CrossRef]
  40. Huang, Y.; Xie, A.; Seidi, F.; Zhu, W.; Li, H.; Yin, S.; Xu, X.; Xiao, H. Core-Shell Heterostructured Nanofibers Consisting of Fe7S8 Nanoparticles Embedded into S-Doped Carbon Nanoshells for Superior Electromagnetic Wave Absorption. Chem. Eng. J. 2021, 423, 130307. [Google Scholar] [CrossRef]
  41. Kuang, D.; Hou, L.; Wang, S.; Yu, B.; Lianwen, D.; Lin, L.; Huang, H.; He, J.; Song, M. Enhanced Electromagnetic Wave Absorption of Ni–C Core-Shell Nanoparticles by HCP-Ni Phase. Mater. Res. Express 2018, 5, 095013. [Google Scholar] [CrossRef]
  42. Yang, R.-X.; Xu, L.-R.; Wu, S.-L.; Chuang, K.-H.; Wey, M.-Y. Ni/SiO2 Core–Shell Catalysts for Catalytic Hydrogen Production from Waste Plastics-Derived Syngas. Int. J. Hydrogen Energy 2017, 42, 11239–11251. [Google Scholar] [CrossRef]
  43. Fu, W.; Yang, H.; Chang, L.; Li, M.; Bala, H.; Yu, Q.; Zou, G. Preparation and Characteristics of Core–Shell Structure Nickel/Silica Nanoparticles. Colloids Surf. A Physicochem. Eng. Asp. 2005, 262, 71–75. [Google Scholar] [CrossRef]
  44. Tang, N.; Lü, L.; Zhong, W.; Au, C.; Du, Y. Synthesis and Magnetic Properties of Carbon-Coated Ni/SiO2 Core/Shell Nanocomposites. Sci. China Ser. G Phys. Mech. Astron. 2009, 52, 31–34. [Google Scholar] [CrossRef]
  45. Xia, L.; Hu, X.; Kang, X.; Zhao, H.; Sun, M.; Cihen, X. A One-Step Facile Synthesis of Ag–Ni Core–Shell Nanoparticles in Water-in-Oil Microemulsions. Colloids Surf. A Physicochem. Eng. Asp. 2010, 367, 96–101. [Google Scholar] [CrossRef]
  46. Zhang, C.; Liu, S.; Mao, Z.; Liang, X.; Chen, B. Ag–Ni Core–Shell Nanowires with Superior Electrocatalytic Activity for Alkaline Hydrogen Evolution Reaction. J. Mater. Chem. A 2017, 5, 16646–16652. [Google Scholar] [CrossRef]
  47. Wang, G.; Wu, H.; Wexler, D.; Liu, H.; Savadogo, O. Ni@Pt Core–Shell Nanoparticles with Enhanced Catalytic Activity for Oxygen Reduction Reaction. J. Alloys Compd. 2010, 503, L1–L4. [Google Scholar] [CrossRef]
  48. Chen, Y.; Gong, C.; Shi, Z.; Chen, D.; Chen, X.; Zhang, Q.; Pang, B.; Feng, J.; Yu, L.; Dong, L. Molten-Salt-Assisted Synthesis of Onion-like Co/CoO@FeNC Materials with Boosting Reversible Oxygen Electrocatalysis for Rechargeable Zn-Air Battery. J. Colloid Interface Sci. 2021, 596, 206–214. [Google Scholar] [CrossRef]
  49. Mazaleyrat, F.; Ammar, M.; LoBue, M.; Bonnet, J.-P.; Audebert, P.; Wang, G.-Y.; Champion, Y.; Hÿtch, M.; Snoeck, E. Silica Coated Nanoparticles: Synthesis, Magnetic Properties and Spin Structure. J. Alloys Compd. 2009, 483, 473–478. [Google Scholar] [CrossRef] [Green Version]
  50. Kim, H.; Achermann, M.; Balet, L.P.; Hollingsworth, J.A.; Klimov, V.I. Synthesis and Characterization of Co/CdSe Core/Shell Nanocomposites: Bifunctional Magnetic-Optical Nanocrystals. J. Am. Chem. Soc. 2005, 127, 544–546. [Google Scholar] [CrossRef]
  51. Liu, Q.; Cao, Q.; Bi, H.; Liang, C.; Yuan, K.; She, W.; Yang, Y.; Che, R. CoNi@SiO2 @TiO2 and CoNi@Air@TiO2 Microspheres with Strong Wideband Microwave Absorption. Adv. Mater. 2016, 28, 486–490. [Google Scholar] [CrossRef]
  52. Bao, F.; Li, J.-F.; Ren, B.; Yao, J.-L.; Gu, R.-A.; Tian, Z.-Q. Synthesis and Characterization of Au@Co and Au@Ni Core−Shell Nanoparticles and Their Applications in Surface-Enhanced Raman Spectroscopy. J. Phys. Chem. C 2008, 112, 345–350. [Google Scholar] [CrossRef]
  53. Lee, W.; Kim, M.G.; Choi, J.; Park, J.-I.; Ko, S.J.; Oh, S.J.; Cheon, J. Redox−Transmetalation Process as a Generalized Synthetic Strategy for Core−Shell Magnetic Nanoparticles. J. Am. Chem. Soc. 2005, 127, 16090–16097. [Google Scholar] [CrossRef]
  54. Lu, Y.; Wang, L.; Chen, M.; Wu, Y.; Liu, G.; Qi, P.; Fu, M.; Wu, H.; Tang, Y. Rationally Designed Hierarchical ZnCo2O4/C Core-Shell Nanowire Arrays for High Performance and Stable Supercapacitors. J. Alloys Compd. 2021, 876, 160037. [Google Scholar] [CrossRef]
  55. Saini, L.; Gupta, V.; Patra, M.K.; Jani, R.K.; Shukla, A.; Kumar, N.; Dixit, A. Impedance Engineered Microwave Absorption Properties of Fe-Ni/C Core-Shell Enabled Rubber Composites for X-Band Stealth Applications. J. Alloys Compd. 2021, 869, 159360. [Google Scholar] [CrossRef]
  56. Nadagouda, M.N.; Varma, R.S. A Greener Synthesis of Core (Fe, Cu)-Shell (Au, Pt, Pd, and Ag) Nanocrystals Using Aqueous Vitamin C. Cryst. Growth Des. 2007, 7, 2582–2587. [Google Scholar] [CrossRef]
  57. Gu, H.; Zheng, R.; Zhang, X.; Xu, B. Facile One-Pot Synthesis of Bifunctional Heterodimers of Nanoparticles: A Conjugate of Quantum Dot and Magnetic Nanoparticles. J. Am. Chem. Soc. 2004, 126, 5664–5665. [Google Scholar] [CrossRef] [PubMed]
  58. Zeng, H.; Li, J.; Wang, Z.L.; Liu, J.P.; Sun, S. Bimagnetic Core/Shell FePt/Fe3O4 Nanoparticles. Nano Lett. 2004, 4, 187–190. [Google Scholar] [CrossRef]
  59. Emadi, M.; Shams, E.; Amini, M.K. Removal of Zinc from Aqueous Solutions by Magnetite Silica Core-Shell Nanoparticles. J. Chem. 2013, 2013, 787682. [Google Scholar] [CrossRef]
  60. Lien, Y.-H.; Wu, T.-M. Preparation and Characterization of Thermosensitive Polymers Grafted onto Silica-Coated Iron Oxide Nanoparticles. J. Colloid Interface Sci. 2008, 326, 517–521. [Google Scholar] [CrossRef] [PubMed]
  61. Santra, S.; Tapec, R.; Theodoropoulou, N.; Dobson, J.; Hebard, A.; Tan, W. Synthesis and Characterization of Silica-Coated Iron Oxide Nanoparticles in Microemulsion: The Effect of Nonionic Surfactants. Langmuir 2001, 17, 2900–2906. [Google Scholar] [CrossRef]
  62. Chen, J.; Wang, Y.G.; Li, Z.Q.; Wang, C.; Li, J.F.; Gu, Y.J. Synthesis and Characterization of Magnetic Nanocomposites with Fe3O4 Core. J. Phys. Conf. Ser. 2009, 152, 012041. [Google Scholar] [CrossRef]
  63. Lee, J.; Lee, Y.; Youn, J.K.; Na, H.B.; Yu, T.; Kim, H.; Lee, S.-M.; Koo, Y.-M.; Kwak, J.H.; Park, H.G.; et al. Simple Synthesis of Functionalized Superparamagnetic Magnetite/Silica Core/Shell Nanoparticles and Their Application as Magnetically Separable High-Performance Biocatalysts. Small 2008, 4, 143–152. [Google Scholar] [CrossRef]
  64. Ta, T.K.H.; Trinh, M.-T.; Long, N.V.; Nguyen, T.T.M.; Nguyen, T.L.T.; Thuoc, T.L.; Phan, B.T.; Mott, D.; Maenosono, S.; Tran-Van, H.; et al. Synthesis and Surface Functionalization of Fe3O4-SiO2 Core-Shell Nanoparticles with 3-Glycidoxypropyltrimethoxysilane and 1,1′-Carbonyldiimidazole for Bio-Applications. Colloids Surf. A Physicochem. Eng. Asp. 2016, 504, 376–383. [Google Scholar] [CrossRef]
  65. Adam, A.; Harlepp, S.; Ghilini, F.; Cotin, G.; Freis, B.; Goetz, J.; Bégin, S.; Tasso, M.; Mertz, D. Core-Shell Iron Oxide@stellate Mesoporous Silica for Combined near-Infrared Photothermia and Drug Delivery: Influence of PH and Surface Chemistry. Colloids Surf. A Physicochem. Eng. Asp. 2022, 640, 128407. [Google Scholar] [CrossRef]
  66. Xuan, S.; Wang, Y.-X.J.; Yu, J.C.; Leung, K.C.-F. Preparation, Characterization, and Catalytic Activity of Core/Shell Fe3O4 @Polyaniline@Au Nanocomposites. Langmuir 2009, 25, 11835–11843. [Google Scholar] [CrossRef] [PubMed]
  67. Nalluri, S.R.; Nagarjuna, R.; Patra, D.; Ganesan, R.; Balaji, G. Large Scale Solid-State Synthesis of Catalytically Active Fe3O4@M (M = Au, Ag and Au-Ag Alloy) Core-Shell Nanostructures. Sci. Rep. 2019, 9, 6603. [Google Scholar] [CrossRef] [PubMed]
  68. Wang, W.; Luo, J.; Fan, Q.; Suzuki, M.; Suzuki, I.S.; Engelhard, M.H.; Lin, Y.; Kim, N.; Wang, J.Q.; Zhong, C.-J. Monodispersed Core−Shell Fe3O4 @Au Nanoparticles. J. Phys. Chem. B 2005, 109, 21593–21601. [Google Scholar] [CrossRef]
  69. Mikhaylova, M.; Kim, D.K.; Bobrysheva, N.; Osmolowsky, M.; Semenov, V.; Tsakalakos, T.; Muhammed, M. Superparamagnetism of Magnetite Nanoparticles: Dependence on Surface Modification. Langmuir 2004, 20, 2472–2477. [Google Scholar] [CrossRef] [PubMed]
  70. Huong, P.T.L.; Huy, L.T.; Lan, H.; Thang, L.H.; An, T.T.; van Quy, N.; Tuan, P.A.; Alonso, J.; Phan, M.-H.; Le, A.-T. Magnetic Iron Oxide-Carbon Nanocomposites: Impacts of Carbon Coating on the As(V) Adsorption and Inductive Heating Responses. J. Alloys Compd. 2018, 739, 139–148. [Google Scholar] [CrossRef]
  71. Li, Z.; Hu, X.; Li, B.; Wang, X.; Shi, Z.; Lu, J.; Wang, Z. MOF-Derived Fe3O4 Hierarchical Nanocomposites Encapsulated by Carbon Shells as High-Performance Anodes for Li-Storage Systems. J. Alloys Compd. 2021, 875, 159906. [Google Scholar] [CrossRef]
  72. Meng, X.; Lei, W.; Yang, W.; Liu, Y.; Yu, Y. Fe3O4 Nanoparticles Coated with Ultra-Thin Carbon Layer for Polarization-Controlled Microwave Absorption Performance. J. Colloid Interface Sci. 2021, 600, 382–389. [Google Scholar] [CrossRef] [PubMed]
  73. Starchikov, S.S.; Zayakhanov, V.A.; Vasiliev, A.L.; Lyubutin, I.S.; Baskakov, A.O.; Nikiforova, Y.; Funtov, K.O.; Lyubutina, M.V.; Kulikova, L.F.; Agafonov, V.N.; et al. Core@shell Nanocomposites Fe7C3/FexOy/C Obtained by High Pressure-High Temperature Treatment of Ferrocene Fe(C5H5)2. Carbon 2021, 178, 708–717. [Google Scholar] [CrossRef]
  74. Yu, G.; Sun, B.; Pei, Y.; Xie, S.; Yan, S.; Qiao, M.; Fan, K.; Zhang, X.; Zong, B. FexOy @C Spheres as an Excellent Catalyst for Fischer−Tropsch Synthesis. J. Am. Chem. Soc. 2010, 132, 935–937. [Google Scholar] [CrossRef]
  75. He, Q.; Zhang, Z.; Xiong, J.; Xiong, Y.; Xiao, H. A Novel Biomaterial—Fe3O4:TiO2 Core-Shell Nano Particle with Magnetic Performance and High Visible Light Photocatalytic Activity. Opt. Mater. 2008, 31, 380–384. [Google Scholar] [CrossRef]
  76. Lu, H.; Yi, G.; Zhao, S.; Chen, D.; Guo, L.-H.; Cheng, J. Synthesis and Characterization of Multi-Functional Nanoparticles Possessing Magnetic, up-Conversion Fluorescence and Bio-Affinity Properties. J. Mater. Chem. 2004, 14, 1336. [Google Scholar] [CrossRef]
  77. Zhao, D.-L.; Teng, P.; Xu, Y.; Xia, Q.-S.; Tang, J.-T. Magnetic and Inductive Heating Properties of Fe3O4/Polyethylene Glycol Composite Nanoparticles with Core–Shell Structure. J. Alloys Compd. 2010, 502, 392–395. [Google Scholar] [CrossRef]
  78. Safaei-Ghomi, J.; Eshteghal, F.; Shahbazi-Alavi, H. A Facile One-Pot Ultrasound Assisted for an Efficient Synthesis of Benzo[g]Chromenes Using Fe3O4/Polyethylene Glycol (PEG) Core/Shell Nanoparticles. Ultrason. Sonochem. 2016, 33, 99–105. [Google Scholar] [CrossRef] [PubMed]
  79. Kim, M.-J.; Choa, Y.-H.; Kim, D.H.; Kim, K.H. Magnetic Behaviors of Surface Modified Superparamagnetic Magnetite Nanoparticles. IEEE Trans. Magn. 2009, 45, 2446–2449. [Google Scholar] [CrossRef]
  80. Chen, F.; Gao, Q.; Hong, G.; Ni, J. Synthesis of Magnetite Core–Shell Nanoparticles by Surface-Initiated Ring-Opening Polymerization of l-Lactide. J. Magn. Magn. Mater. 2008, 320, 1921–1927. [Google Scholar] [CrossRef]
  81. Sakaguchi, M.; Makino, M.; Ohura, T.; Yamamoto, K.; Enomoto, Y.; Takase, H. Surface Modification of Fe3O4 Nanoparticles with Dextran via a Coupling Reaction between Naked Fe3O4 Mechano-Cation and Naked Dextran Mechano-Anion: A New Mechanism of Covalent Bond Formation. Adv. Powder Technol. 2019, 30, 795–806. [Google Scholar] [CrossRef]
  82. Qin, S.; Wang, L.; Zhang, X.; Su, G. Grafting Poly(Ethylene Glycol) Monomethacrylate onto Fe3O4 Nanoparticles to Resist Nonspecific Protein Adsorption. Appl. Surf. Sci. 2010, 257, 731–735. [Google Scholar] [CrossRef]
  83. Omer-Mizrahi, M.; Margel, S. Synthesis and Characterization of Magnetic and Non-Magnetic Core–Shell Polyepoxide Micrometer-Sized Particles of Narrow Size Distribution. J. Colloid Interface Sci. 2009, 329, 228–234. [Google Scholar] [CrossRef]
  84. Yuan, W.; Yuan, J.; Zhou, L.; Wu, S.; Hong, X. Fe3O4@poly(2-Hydroxyethyl Methacrylate)-Graft-Poly(ɛ-Caprolactone) Magnetic Nanoparticles with Branched Brush Polymeric Shell. Polymer 2010, 51, 2540–2547. [Google Scholar] [CrossRef]
  85. Chen, S.; Li, Y.; Guo, C.; Wang, J.; Ma, J.; Liang, X.; Yang, L.-R.; Liu, H.-Z. Temperature-Responsive Magnetite/PEO−PPO−PEO Block Copolymer Nanoparticles for Controlled Drug Targeting Delivery. Langmuir 2007, 23, 12669–12676. [Google Scholar] [CrossRef]
  86. Qin, R.; Li, F.; Jiang, W.; Chen, M. Synthesis and Characterization of Diethylenetriaminepentaacetic Acid-Chitosan-Coated Cobalt Ferrite Core/Shell Nanostructures. Mater. Chem. Phys. 2010, 122, 498–501. [Google Scholar] [CrossRef]
  87. Tang, X.; Jia, R.; Zhai, T.; Xia, H. Hierarchical Fe3O4 @Fe2O3 Core–Shell Nanorod Arrays as High-Performance Anodes for Asymmetric Supercapacitors. ACS Appl. Mater. Interfaces 2015, 7, 27518–27525. [Google Scholar] [CrossRef] [PubMed]
  88. Carnes, C.L.; Klabunde, K.J. Unique Chemical Reactivities of Nanocrystalline Metal Oxides toward Hydrogen Sulfide. Chem. Mater. 2002, 14, 1806–1811. [Google Scholar] [CrossRef]
  89. Shen, H.; Yao, J.; Gu, R. Fabrication and Characteristics of Spindle Fe2O3@Au Core/Shell Particles. Trans. Nonferrous Met. Soc. China 2009, 19, 652–656. [Google Scholar] [CrossRef]
  90. Utech, S.; Scherer, C.; Krohne, K.; Carrella, L.; Rentschler, E.; Gasi, T.; Ksenofontov, V.; Felser, C.; Maskos, M. Magnetic Polyorganosiloxane Core–Shell Nanoparticles: Synthesis, Characterization and Magnetic Fractionation. J. Magn. Magn. Mater. 2010, 322, 3519–3526. [Google Scholar] [CrossRef]
  91. Galperin, A.; Margel, S. Synthesis and Characterization of Radiopaque Magnetic Core-Shell Nanoparticles for X-ray Imaging Applications. J. Biomed. Mater. Res. Part B Appl. Biomater. 2007, 83B, 490–498. [Google Scholar] [CrossRef] [PubMed]
  92. Thünemann, A.F.; Schütt, D.; Kaufner, L.; Pison, U.; Möhwald, H. Maghemite Nanoparticles Protectively Coated with Poly(Ethylene Imine) and Poly(Ethylene Oxide)- b Lock -Poly(Glutamic Acid). Langmuir 2006, 22, 2351–2357. [Google Scholar] [CrossRef]
  93. Nagao, D.; Yokoyama, M.; Yamauchi, N.; Matsumoto, H.; Kobayashi, Y.; Konno, M. Synthesis of Highly Monodisperse Particles Composed of a Magnetic Core and Fluorescent Shell. Langmuir 2008, 24, 9804–9808. [Google Scholar] [CrossRef]
  94. Kaur, P.; Thakur, R.; Malwal, H.; Manuja, A.; Chaudhury, A. Biosynthesis of Biocompatible and Recyclable Silver/Iron and Gold/Iron Core-Shell Nanoparticles for Water Purification Technology. Biocatal. Agric. Biotechnol. 2018, 14, 189–197. [Google Scholar] [CrossRef]
  95. Lehner, R.; Wang, X.; Marsch, S.; Hunziker, P. Intelligent Nanomaterials for Medicine: Carrier Platforms and Targeting Strategies in the Context of Clinical Application. Nanomed. Nanotechnol. Biol. Med. 2013, 9, 742–757. [Google Scholar] [CrossRef]
  96. Platania, V.; Kaldeli-Kerou, A.; Karamanidou, T.; Kouki, M.; Tsouknidas, A.; Chatzinikolaidou, M. Antibacterial Effect of Colloidal Suspensions Varying in Silver Nanoparticles and Ions Concentrations. Nanomaterials 2021, 12, 31. [Google Scholar] [CrossRef] [PubMed]
  97. Ntasiou, P.; Kaldeli Kerou, A.; Karamanidou, T.; Vlachou, A.; Tziros, G.T.; Tsouknidas, A.; Karaoglanidis, G.S. Synthesis and Characterization of Novel Copper Nanoparticles for the Control of Leaf Spot and Anthracnose Diseases of Olive. Nanomaterials 2021, 11, 1667. [Google Scholar] [CrossRef] [PubMed]
  98. Machado, N.; Bruininks, B.M.H.; Singh, P.; dos Santos, L.; Dal Pizzol, C.; Dieamant, G.d.C.; Kruger, O.; Martin, A.A.; Marrink, S.J.; Souza, P.C.T.; et al. Complex Nanoemulsion for Vitamin Delivery: Droplet Organization and Interaction with Skin Membranes. Nanoscale 2022, 14, 506–514. [Google Scholar] [CrossRef] [PubMed]
  99. Martin, C.R. Nanomaterials: A Membrane-Based Synthetic Approach. Science 1994, 266, 1961–1966. [Google Scholar] [CrossRef] [PubMed]
  100. Karamanidou, T.; Tsouknidas, A. Plant-Derived Extracellular Vesicles as Therapeutic Nanocarriers. Int. J. Mol. Sci. 2021, 23, 191. [Google Scholar] [CrossRef] [PubMed]
  101. Cammas, S.; Suzuki, K.; Sone, C.; Sakurai, Y.; Kataoka, K.; Okano, T. Thermo-Responsive Polymer Nanoparticles with a Core-Shell Micelle Structure as Site-Specific Drug Carriers. J. Control. Release 1997, 48, 157–164. [Google Scholar] [CrossRef]
  102. Law, W.-C.; Yong, K.-T.; Roy, I.; Xu, G.; Ding, H.; Bergey, E.J.; Zeng, H.; Prasad, P.N. Optically and Magnetically Doped Organically Modified Silica Nanoparticles as Efficient Magnetically Guided Biomarkers for Two-Photon Imaging of Live Cancer Cells. J. Phys. Chem. C 2008, 112, 7972–7977. [Google Scholar] [CrossRef]
  103. Kateb, B.; Chiu, K.; Black, K.L.; Yamamoto, V.; Khalsa, B.; Ljubimova, J.Y.; Ding, H.; Patil, R.; Portilla-Arias, J.A.; Modo, M.; et al. Nanoplatforms for Constructing New Approaches to Cancer Treatment, Imaging, and Drug Delivery: What Should Be the Policy? NeuroImage 2011, 54, S106–S124. [Google Scholar] [CrossRef] [Green Version]
  104. Kayal, S.; Ramanujan, R.V. Anti-Cancer Drug Loaded Iron–Gold Core–Shell Nanoparticles (Fe@Au) for Magnetic Drug Targeting. J. Nanosci. Nanotechnol. 2010, 10, 5527–5539. [Google Scholar] [CrossRef]
  105. Berry, C.C.; Curtis, A.S.G. Functionalisation of Magnetic Nanoparticles for Applications in Biomedicine. J. Phys. D Appl. Phys. 2003, 36, R198–R206. [Google Scholar] [CrossRef]
  106. Hegyi, G.; Szigeti, G.P.; Szász, A. Hyperthermia versus Oncothermia: Cellular Effects in Complementary Cancer Therapy. Evid.-Based Complement. Altern. Med. 2013, 2013, 672873. [Google Scholar] [CrossRef]
  107. Gilchrist, R.K.; Medal, R.; Shorey, W.D.; Hanselman, R.C.; Parrott, J.C.; Taylor, C.B. Selective Inductive Heating of Lymph Nodes. Ann. Surg. 1957, 146, 596–606. [Google Scholar] [CrossRef] [PubMed]
  108. Tajabadi, M.; Rahmani, I.; Mirkazemi, S.M.; Goran Orimi, H. Insights into the Synthesis Optimization of Fe@SiO2 Core-Shell Nanostructure as a Highly Efficient Nano-Heater for Magnetic Hyperthermia Treatment. Adv. Powder Technol. 2022, 33, 103366. [Google Scholar] [CrossRef]
  109. Chen, Z.; Song, S.; Ma, J.; da Ling, S.; Wang, Y.D.; Kong, T.T.; Xu, J.H. Fabrication of Magnetic Core/Shell Hydrogels via Microfluidics for Controlled Drug Delivery. Chem. Eng. Sci. 2022, 248, 117216. [Google Scholar] [CrossRef]
  110. Etemadi, H.; Plieger, P.G. Magnetic Fluid Hyperthermia Based on Magnetic Nanoparticles: Physical Characteristics, Historical Perspective, Clinical Trials, Technological Challenges, and Recent Advances. Adv. Ther. 2020, 3, 2000061. [Google Scholar] [CrossRef]
Figure 1. Top-down and bottom-up processes.
Figure 1. Top-down and bottom-up processes.
Metals 12 00605 g001
Figure 2. High-energy ball milling process.
Figure 2. High-energy ball milling process.
Metals 12 00605 g002
Figure 3. From left to right: photolithography and electron beam lithography.
Figure 3. From left to right: photolithography and electron beam lithography.
Metals 12 00605 g003
Figure 4. The sol-gel process.
Figure 4. The sol-gel process.
Metals 12 00605 g004
Figure 5. Micelle (left) and reverse micelle (right).
Figure 5. Micelle (left) and reverse micelle (right).
Metals 12 00605 g005
Figure 6. Teflon-lined stainless-steel autoclave used for hydrothermal and solvothermal reactions.
Figure 6. Teflon-lined stainless-steel autoclave used for hydrothermal and solvothermal reactions.
Metals 12 00605 g006
Figure 7. Hot-injection flask.
Figure 7. Hot-injection flask.
Metals 12 00605 g007
Figure 8. Hot-wall and cold-wall CVD reactors.
Figure 8. Hot-wall and cold-wall CVD reactors.
Metals 12 00605 g008
Figure 9. Core-shell nanoparticles for combined drug delivery and hyperthermia applications.
Figure 9. Core-shell nanoparticles for combined drug delivery and hyperthermia applications.
Metals 12 00605 g009
Table 1. Overview of production methods for magnetic core/shell nanoparticles.
Table 1. Overview of production methods for magnetic core/shell nanoparticles.
CoreShell
Core/ShellShapeAverage SizeMethodBasic ReagentMethodBasic ReagentRef.
Fe/Cirregular spherical16–41 nmthermal decompositionFeCl3·6H2O fructose, glucose, and sucrose[26]
Fe/Cspherical25 nmDC plasma arc discharge methodFe powder (high purity) carbon powder[27]
Fe/Cspherical200 nmreduction with CVDFe3O4 carbon powder[28]
Fe/Cspherical50–60 nmthermal decompositionFerrocene, silicon waferthermal decompositionheavy oil[29]
Fe/SiO2spherical agglomerates97 nmgas-phase hybridizationFe powdergas-phase hybridizationSiO2 powder[30]
Fe/SiO2spherical10–40 nmreduction by KBH4FeCl2·4H2Ohydrolysis, condensation polymerizationTEOS[31]
Fe/SiO2spherical10–90 nmarc dischargeFe powder SiO2 powder[32]
Fe/FeO, Fe3O4spherical30–80 nmsolar vapor phase condensationFe powdersolar vapor phase condensationFe powder,
Fe3O4 powder
[33]
Fe/Agspherical33 nmreduction by NaBH4FeSO4·7H2O AgNO3[34]
Fe/Agspherical10 nmreduction by LiBEt3HFeCl2reduction in organic mediaAgNO3[35]
Ag/Fespherical30 nmreduction via phytochemicalsFe(NO3)3 AgNO3[36]
Fe/Ptspherical≈2 nmreduction by NaBH4FeSO4reduction transmetalationK2PtCl4[37]
Fe/Ptspherical3 nmreductionC15H21FeO6reductionPtCl4[38]
Fe/PIBspherical20–100 nmthermal decompositionFe(CO)5 PIB[39]
Fe/PSspherical20–100 nmthermal decompositionFe(CO)5 polystyrene (PS)[39]
Fe-CBC/SdCnanoparticles on fibers<100 nm fiber diameter (NPs << 50 nm)cellulose-assisted hydrolysis, polymer wrappingFeCl3, BC hydrogelscarbonizationPEDOT, EDOT[40]
Ni/Cspherical14 nmreduction with CVDNi(acac)2annealing [41]
Ni/C 15–35 nmthermal decompositionNi(C5H7O2)2thermal decompositionNi(C5H7O2)2[29]
SiO2/Nispherical150 nmStöber methodTEOSdeposition, precipitationNi(NO3)2·6H2O[42]
Ni/SiO2spherical80 nmWEE techniqueNi wireStöber methodTEOS[43]
Ni/SiO2elliptic and spherical≈300 nm (20–80 nm shell)reduction by citric acidNiCl2·6H2OStöber methodTEOS[44]
Ag/Nispherical50–100 nm
95–165 nm
reduction by NaBH4 in microemulsionNi(NO3)2reduction by NaBH4 in microemulsionAgNO3[45]
Ag/Ninanowire (and spherical)<300 nm
(50–150 nm shell)
hydrothermal methodAgNO3hydrothermal methodNi(AC)2[46]
Ni/Ptirregular spherical agglomerated<10 nmreduction by NaBH4Ni(CH3COO)2reduction by NaBH4
and N2H4
H2PtCl4·6H2O[47]
Co/CoO@FeNCspherical core onion-like shells50 nmmicroemulsion, pyrolysisFeCl3, CoCl2·6H2O, RuO2, Pt/C, cyclohexane [48]
Co/SiO2spherical≈60 nmcryogenic
melting method
Co bulk metalStöber methodTEOS[49]
Co/CdSespherical≈15 nmhigh-temperature thermal decompositionCo2(CO)8precipitationCd(CH3)2, Se[50]
CoNi/SiO2/TiO2spherical1.35 µm/100 nm shellsolvothermalC4H6NiO4 4H2O,
C4H6CoO4 4H2O
Stöber method,
solvothermal
TEOS,
TIP
[51]
CoNi@Air@TiO2Spherical yolk type1.35 µm/200 nm shellsolvothermalC4H6NiO4 4H2O,
C4H6CoO4 4H2O
Stöber method, solvothermal, hydrothermal via NaOHTEOS, TIP[51]
Au/Cospherical65 nmreduction by sodium citrateHAuCl4reduction by N2H4, H2OCoCl2[52]
Au/Nispherical65 nmreduction by sodium citrateHAuCl4reduction by N2H4, H2ONiCl2[52]
Co/Au, Pd, Pt, Cuspherical≈6.5 nmHydrothermal decompositionCo2(CO)8Reduction transmetalationAu precursor Pd(hfac)2, Pt(hfac)2, Cu(hfac)2[53]
ZnCo2O4/Cnanowire arrays100 nmhydrothermal, annealingZn(NO3)2·6H2O,
Co(NO3)2·6H2O, urea
CVDC2H2[54]
FeNi/Cspherical50 nmprecipitation, reduction with NaOHFeCl2, NiCl2, aniline, hydrochloric acid, formaldehydepyrolysis [55]
FeNi/SiO2spherical≈60 nmcryogenic melting methodFe, Ni bulk metalStöber methodTEOS[49]
Fe, Cu/
Au, Pt, Pd, Ag
cubical, spherical5–50 nm, 50–60 nmreduction by vitamin CFe(NO3)3 9H2O, CuCl2reduction by vitamin CNa2PtCl6·6H2O, HAuCl4·3H2O,
AgNO3, PdCl2
[56]
FePt/CdSspherical<10 nmthermal decomposition with reductionFe(CO)5, Pt(C5H7O2)2precipitationsulfur, Cd(C5H7O2)2[57]
Fe58Pt42/Fe3O4spherical4–7 nmreduction with thermal decompositionPt(C5H7O2)2, Fe(CO)5thermal decomposition1,2-hexadecanediol, oleic acid, Fe(C5H7O2)3 , oleylamine[58]
MnFe2O4/TEHA-co-PDLLA @PTXspherical, irregularly shaped particles<160 nmsolvothermalacetylacetonate iron (III), manganese (II), Fe(acac)3, Mn(acac)2emulsionTEHA-co-PDLLA block copolymer, Paclitaxel[2]
Fe3O4/SiO2spherical30 nmco-precipitationFeCl2⋅4H2O, FeCl3⋅6H2OStöber methodTEOS[59]
Fe3O4/SiO2spherical30 nmthermal decompositionFe(C5H7O2)3sol-gelTEOS[60]
Fe3O4/SiO2acicular<5 nmchemical reaction in microemulsionFeCl3, FeSO4sol-gel reaction in microemulsionTEOS[61]
Fe3O4/SiO2spherical15–20 nmwet chemical reactionFeCl3, FeSO4hydrolysisNa2SiO3[62]
Fe3O4/SiO2spherical7–12 nmwet chemical reactionFeCl3, N2H4sol-gelTEOS[63]
Fe3O4/SiO2spherical80 nmco-precipitationFeCl2·4H2O, FeCl3·6H2Osonication, hydrolysisTEOS[64]
Fe3O4/SiO2stellate100 nmthermal decompositionFeSt3sol-gelTEOS, CTATos[65]
Fe3O4/Polyaniline/Auirregular spherical200 nm (4 nm Au NPs as shell)solvothermalFeCl3·6H2O, sodium acrylate, CH3COONareduction by
Na3C6H5O7, NaBH4
HAuCl4[66]
Fe3O4/Auoctahedral100–150 nmphysical grindingcommercial Fe3O4calcinationHAuCl4[67]
Fe3O4/Auspherical7 nmsolvothermalFe(C5H7O2)3, phenyl ether, oleic acid, oleylaminehydrolysis with thermal decompositionAu(OOCCH3)3, 1,2-hexadecanediol, oleic acid, oleylamine[68]
Fe3O4/Auspherical8 nmwet chemical reactionFeCl3reduction by NaBH4HAuCl4[69]
Fe3O4/Agoctahedral100–150 nmphysical grindingcommercial Fe3O4calcinationAgNO3[67]
Fe3O4/Cspherical~20 nmco-precipitationFeCl3·6H2O, FeCl2·4H2Ohydrothermalglucose[70]
Fe3O4/Crod, diamond, spindle2 μm, 3 μm, 600 nmMOF-derived method and solvothermalFeCl3·6H2O, fumaric acid, dimethyl formamidepyrolysis [71]
Fe3O4/Cspherical52, 80, and 110 nmHigh-temperature solution-phase reactionFe(acac)3, oleic acid, dibenzyl ether, 1-Octadecene, hexanecarbonization [72]
Fe7C3/FexOy/Cspherical12–45 nmthermal decompositionFe(C5H5)2thermal decomposition [73]
FexOy/Cspherical7 nm @6 um sphereswet chemical reactionFe(NO3)3·6H2Othermal decompositionglucose[74]
Fe3O4/TiO2spherical20 nmwet chemical reactionFeCl3·6H2O, FeCl2·4H2OprecipitationTi(SO4)2, CO(NH2)2[75]
Fe3O4/SiO2/TiO2irregular spherical≈30 nmwet chemical reactionFeCl3, FeSO4sol-gelTEOS, TBOT[62]
Fe3O4/SiO2/Al2O3irregular spherical20–25 nmwet chemical reactionFeCl3, FeSO4sol-gel with precipitationTEOS[62]
Fe3O4, γ-Fe2O3/NaYF4:Yb, Erspherical68 nmco-precipitationFeCl2, FeCl3co-precipitationNaF, YCl3, YbCl3, ErCl3, EDTA[76]
Fe3O4/PEGspherical10–40 nmwet chemical reactionFeCl2·4H2O, FeCl3·6H2O PEG-400[77]
Fe3O4/PEGspherical2–10 nmco-precipitationFeCl2·4H2O, FeCl3·6H2Oco-precipitation, sonicationPEG-400[78]
Fe3O4/PEGirregular spherical12 nmco-precipitationFeCl3 PEG (mol wt. 4000)[79]
Fe3O4/PLAspherical10 nmco-precipitationFeCl3·6H2O, FeCl2·4H2Oring-opening
polymerization
L-lactide[80]
Fe3O4/Dextranspherical≈35 nm Fe3O4 powdervibration ball millingDextran 60000[81]
Fe3O4/MPEGspherical7.8 nmwet chemical reactionFeCl3·6H2O, FeCl2·4H2O MPEG (mol wt. 5000)[69]
Fe3O4/PEGMAquasi spherical15 nmco-precipitationFeCl3·6H2O, FeCl2·4H2ORAFTPEGMA[82]
PEGMA/PS/Fe3O4quasi spherical3.7 um (90 nm Fe3O4 NPs)polymerizationGMA, EDMAco-precipitationFeCl3, NaNO2[83]
Fe3O4/PHEMA-g-PCLspherical4 nmHydrothermal degradationFe(OCOCH3)3polymerizationpentamethyldiethylenetriamine[84]
Fe3O4/PEO-PPO-PEOspherical45–20 nmreductionFeCl3 block polymer
PEO-PPO-PEO
[85]
CoFe2O4/DTPA-CSspherical70 nmlow temperature solid-state methodCoCl2·6H2O, FeCl3·6H2O, NaClemulsion cross-linking
polymerization
chitosan, DTPA[86]
Fe3O4/Fe2O3nanorod30−50 nm (diameter), 0.8−1.2 μm (length)hydrothermalFeCl3·6H2O, Na2SO4electrodepositionFeOOH, FeCl2[87]
CaO/Fe2O3irregular9–16 nm bulk CaOthermal decompositionFe(C5H7O2)3[88]
MgO/Fe2O3irregular6–9 nmaerogel/hypercritical drying/ dehydrationMg(OCH3)2thermal decompositionFe(C5H7O2)3[88]
Fe2O3/Auspindle80 nm × 500 nmhydrolysis in KH2PO4 solutionFeCl3 6H2Oreduction by
formaldehyde
HAuCl4[89]
Fe2O3/POSspherical15–26 nmco-precipitationFeCl2, FeCl3polycondensationTMMS, DEDMS, ClBz-T[90]
γ-Fe2O3 /polyMAOETIBspherical≈56 nmprecipitationgelatin, iron oxideemulsion polymerization2-MAOETIB[91]
γ-Fe2O3/PEI+PEO-PGAIrregular spherical≈45 nmco-precipitationFeCl2, FeCl3suspension polymerizationPEO-PGA[92]
iron oxide-SiO2 composite/PSspherical≈268 nmMassart with Stöber methodFeCl2, FeCl3, TEOSpolymerizationstyrene[93]
FeO/Agspherical20–30 nmwet chemical reaction with peel extract (PEP)FeCl3 AgNO3[94]
FeO/Auspherical<100 nmwet chemical reaction with peel extract (PEP)FeCl3 AuCl3[94]
Abbreviations: CVD = Chemical Vapor Deposition; TEOS = tetraethoxysilane; PIB = polyisobutylene; GMA = glycidyl methacrylate; EDMA = ethylene glycol dimethacrylate; PEO-PPO-PEO = poly(ethylene oxide)-poly(propylene oxide)-poly-(ethylene oxide); DTPA = diethylenetriaminepentaacetic acid; TMMS = trimethoxymethylsilane; DEDMS = diethoxydimethylsilane; ClBz-T = (chloromethylphenyl)trimethoxysilane; 2-MAOETIB = 2-methacryloyloxyethyl(2,3, 5-triiodobenzoate); PTX = Paclitaxel; TEHA-co-PDLLA = thioether-containing ω-hydroxyacid-co-poly(d,l-lactic acid); CBC =carbonized bacterial cellulose; SdC = S-doped carbon; EDOT = 3, 4-ethoxylene dioxy thiophene; hfac = 1,1,1,5,5,5-hexafluoroacetylacetonat; CTATos = cetyltrimethylammonium tosylate; PHEMA-g-PCL = poly (2-hydroxyethyl methacrylate)-graft- poly(e-caprolactone); POS = polyorganosiloxane.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Tsamos, D.; Krestou, A.; Papagiannaki, M.; Maropoulos, S. An Overview of the Production of Magnetic Core-Shell Nanoparticles and Their Biomedical Applications. Metals 2022, 12, 605. https://doi.org/10.3390/met12040605

AMA Style

Tsamos D, Krestou A, Papagiannaki M, Maropoulos S. An Overview of the Production of Magnetic Core-Shell Nanoparticles and Their Biomedical Applications. Metals. 2022; 12(4):605. https://doi.org/10.3390/met12040605

Chicago/Turabian Style

Tsamos, Dimitris, Athina Krestou, Maria Papagiannaki, and Stergios Maropoulos. 2022. "An Overview of the Production of Magnetic Core-Shell Nanoparticles and Their Biomedical Applications" Metals 12, no. 4: 605. https://doi.org/10.3390/met12040605

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