Next Article in Journal
Interrelation between the Solid-State Synthesis Conditions and Magnetic Properties of the NiCr2O4 Spinel
Next Article in Special Issue
Magnetism and EPR Spectroscopy of Nanocrystalline and Amorphous TiO2: Fe upon Al Doping
Previous Article in Journal
Up-Conversion Luminescence and Magnetic Properties of Multifunctional Er3+/Yb3+-Doped SiO2-GdF3/LiGdF4 Glass Ceramics
Previous Article in Special Issue
Magnetic Nylon 6 Nanocomposites for the Microextraction of Nucleic Acids from Biological Samples
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Magnetic Nanocomposites and Imprinted Polymers for Biomedical Applications of Nucleic Acids

Institute of Chemical Biology and Fundamental Medicine SB RAS, 630090 Novosibirsk, Russia
*
Author to whom correspondence should be addressed.
Magnetochemistry 2023, 9(1), 12; https://doi.org/10.3390/magnetochemistry9010012
Submission received: 22 November 2022 / Revised: 15 December 2022 / Accepted: 28 December 2022 / Published: 30 December 2022

Abstract

:
Magnetic nanocomposites (MNCs) combine the features of magnetic nanoparticles and a second material, which provide distinct physical, chemical, and biological properties. The magnetic core for nanocomposite synthesis is extensively used due to its high saturation magnetization, chemical stability, large surface area, and easy functionalization. Moreover, magnetic nanoparticles (MNPs) have great potential for magnetic resonance imaging (MRI), magnetic particle imaging (MPI), hyperthermia, and targeted drug and gene delivery by an external magnetic field. Numerous composing units exist, which leads to the outstanding application of composites. This review focuses on nucleic acid-based bioapplications of MNCs with polymeric, organic, inorganic, biomolecules, and bioinspared surface coating. In addition, different forms, such as core–shell, doping, multilayer, yolk–shell, and Janus-shaped hybrids, are discussed, and their unique properties are highlighted. The unique types of nanocomposites as magnetic molecularly imprinted polymer (MMIP) properties are presented. This review presents only the synthesis of MNCs using ready-made magnetic cores. These restrictions are associated with many materials, the quantitative and qualitative magnetic core composition, and synthesis procedures. This review aims to discuss the features of nucleic acid-based MNC information available to researchers in this field and guide them through some problems in the area, structure variation, and surface functionalization possibilities. The most recent advancements of MNCs and imprinted polymers in nucleic acid-based therapy, diagnostics, theranostics, magnetic separation, biocatalytic, and biosensing are introduced.

Graphical Abstract

1. Introduction

Magnetic nanocomposites (MNCs) are multiphase solid materials where one phase of a nanoscale material is magnetic, and another is varying. Such a combination allows for the creation of new materials with different properties from the initial nanoparticles. Various building blocks of MNCs may be small molecules (tannic, lauric, myristic, oleic acids, etc.) [1,2,3,4], organic polymers (polyethylene glycol, polyethylene imine, tween, etc.) [5,6,7,8,9,10,11], inorganic metals (gold, platinum, etc.) [12,13,14,15], oxides (silica) [16,17,18,19], salts (calcium carbonate) [20,21], or bioinspired materials (proteins, carbohydrates, nucleic acid, aptamers, polydopamine) [5,8,22,23,24,25,26,27,28] (Figure 1). A number of reviews highlight the possible features of such a surface stabilization and functionalization [12,22,25,29,30,31]. The magnetic core of MNCs has become vital for a wide range of applications, including material science, hyperthermia, the contrast agent area for magnetic resonance imaging (MRI) and magnetic particle imaging (MPI), and theranostics [32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48]. Manipulation with an external magnetic field is essential for magnetic separation, drug and gene delivery, biomass processing, and biosensing [39,42,49,50,51].
The first reported iron nanoparticles were synthesized in 1984 [52]. However, about thirty years were required to be able to synthesize stable magnetic nanoparticles (MNPs) with optimal physical properties. One of the most common MNPs for biomedical applications is magnetite (Fe3O4) due to its ferrimagnetism, high stability, and cost-effectivity [31]. Many papers were produced which showed the use of magnetite MNPs for different applications [30,39,40,43,53,54,55,56,57].
One of the milestones was the synthesis of silica-coated MNPs in 1995, which opened the era of MNCs. Silica-coated and functionalized MNCs represent a significant doorway for nucleic acid (NA) separation, purification, and detection [50,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74]. NA is one of the most important substances in cells and has an enormous number of biological functions. Functionally integrating NA and MNPs produces a rich variety of MNCs which, in many cases, display unique or augmented properties due to the synergistic effect. The NA component may be in various forms, such as plasmid DNA, single-stranded (ssNA) or double-stranded (dsNA) DNA or RNA, small interfering RNA (siRNA), micro-RNA (miRNA), short hairpin RNA (shRNA), antisense oligonucleotides, etc. [75,76]. NA may be adsorbed on the MNCs or chemically bound to the surface, which requires preliminary modification. Thus, the design of NA-based MNCs is a challenging task and depends on the ultimate purpose. The NA component typically provides sequence-based addressability for probes, and complex spatial architecture (e.g., DNA origami, aptamers, G-quadruplex, triplex form, etc.) for specific binding with biopolymers, biological selectivity, or disease treatment [63,70,74]. Numerous NA-based MNCs have been developed for disease analysis purposes, which exhibit more advantages than traditional systems [63,69,74]. MNCs containing gene vectors possess unique advantages for their rapid transfection, isolation, and elevated gene shuffling into the cells [77,78,79,80]. The NA-based MNC tools provide diverse applications, which include targeted delivery, therapy, imaging, biosensing, and diagnostics [59,63,69,70,72,74,78,81,82,83]. This review highlights only the synthesis of MNCs using ready-made magnetic cores. These restrictions are associated with many materials, the quantitative and qualitative magnetic core composition, and synthesis procedures. This review is divided into two main sections. Each section provides a brief overview of some relevant NA-based MNC properties. The first section describes the structural types of MNCs with organic, inorganic, and biomolecule coatings. The second section is focused on NA-based MNC applications, highlighting multifunctional and bioinspired materials, probes, smart constructions, and devices. Some key issues of an MNC structure and the importance of advanced materials for various applications are proposed. The outlook on the future directions and challenges are discussed.

2. Magnetic Nanocomposite Types

MNCs are multiphase systems that can combine the properties of their component materials. Many research articles focused on the synthesis and MNP coating procedures to obtain the desired morphology, particle sizes, and physicochemical properties [43,53]. The surfaces of MNPs used for biomedical applications are usually modified with various coatings and biological molecules to acquire a good aqueous stability, colloid stability, low toxicity, biocompatibility, and recognition by tissues or cells. Small organic molecules [1,2,3,4], artificial and natural polymers [5,8,22,23,24,25,26], and inorganic substances [12,13,14,15,16,17,18,19] are used for coating procedures [84,85,86,87,88].
The main structural types of nanocomposites are divided into core–shell [89,90,91], Janus [92,93,94], assembly [95,96], yolk–shell [97,98], multicomponent [99], multilayer shell [100], etc. (Figure 2) [101,102]. The core–shell structural type is generally biocompatible and is well tolerated in vivo [44]. The non-toxic properties appear due to a good magnetic core surface protection and low interaction with a solvent. The core–shell MNC usually has a well-controlled structure and tunable physicochemical properties, depending on the surfactant, coating type, and biocompatibility, and possible surface functionalization [89,90,91,102]. They primarily do not have a problem with magnetic core toxicity but may have with surfactant polymer biocompatibility. For example, PEI coating exhibited a significantly higher cell uptake but may cause severe cytotoxicity through multiple mechanisms. However, replacing PEI with a biocompatible coating may easily solve the problem [44]. Nowadays, core–shell MNCs are used in numerous potential issues such as biocatalysis, magnetic separation, biosensing, and nanosorbents [90,102]. Nevertheless, the synthesis procedure is not simple to yield a monodisperse and controlled shell thickness [90]. Yolk–shell nanocomposites are a subtype of the core–shell structure. The only difference is an interior void is located between the core and the shell [97,98,102]. An example of such a structural type is an MNP coated with a silica interlayer and porous gold outer shell, followed by removing the silica template [102]. The yolk–shell structure’s extra feature is the diffusion possibility into and out of the shell. The cavity of the yolk–shell is a host place for drugs, adsorbed molecules, and catalytic activity [97,98,102]. Janus particle is a type that is divided into two parts, one of which is a repulsive core and the other is a highly interactive surface. The resulting MNC combines two sets of properties (magnetic core and various coating), which, unlike the core–shell composites, possess two types of properties. This technology may be useful for biosensors, protonics, imaging, and nanomedicine applications [93,94,97]. This section is divided into three subsections that highlight organic molecules, polymers, biomolecules, inorganic compounds, and other types of MNP coatings.

2.1. Organic Molecules and Polymers-Coated MNPs

MNPs have a high surface energy and become unstable, as well as having a tendency to aggregate. The properties of MNPs’ surface critically influence the overall performance of the material. The small organic molecules and a polymer coating may shield the MNPs’ core from oxidation by the harsh chemical environment. Moreover, this enhances the colloidal stability of nanoparticles and their biocompatibility in the physiological media [22,29]. Small organic molecules or a polymer coating may be hydrophilic, hydrophobic, amphiphilic, or have a charge which highly restricts further application. A modern approach uses fatty or amphiphilic compounds, such as carboxylic acids (polyacrylic, lauric, myristic, or oleic) [103], alkylsulfonic acids, alkylphosphonic acids, polyphenols (tannic acid), alcohols or ethers (polyethylene glycol (PEG), polysorbate (Tween), polyvinylethanol) [5,104], carbohydrates or their derivatives (hyaluronic acid, dextran, chitosan) [5], amines (polyethyleneimine, PEI), polyamides (nylon 6), etc. [2,3,4,6,7,8,9,10,11,19,25,26,83,88,102,105,106,107,108,109,110,111,112,113,114,115]. Especially, polymers represent an excellent class of compounds for the high-capacity binding of biopolymers such as nucleic acids, peptides, enzymes, proteins, and lipids [112]. Structure variability and functional groups of polymers make the great diversification of MNCs’ properties possible. The charge, solubility in water or organic solvents, viscosity, film-forming ability, pH-, ionic strength, and temperature stability may be changed. In addition, polymer-coated MNCs usually have a little tendency to aggregate and a higher colloidal stability, which significantly increases their applicability (Figure 3). For example, one of the most common polymeric ligands, PEG, usually increases the colloidal stability and solubility in water and removes non-specific protein adsorption and phagocyte uptake [7,88,116,117,118,119,120]. Tween family surfactants consist of a hydrophilic head group and a different-length hydrophobic alkyl chain [7,121]. Therefore, the desired hydrophile–hydrophobic tween molecule may be used for coating. However, the organic functionality is typically used to make the nanoparticles hydrophilic and biocompatible and to endow their colloidal stability in water. Small organic molecules and polymers allow simple nucleic acids and their derivatives’ adsorption on the surface due to the positive charge and hydrophobic–hydrophilic balance [61,70,72,102,120,122]. The MNCs provide a simple and quick nucleic acid separation from the supernatant by applying a magnet.
PEI is one of the widely used positive-charged polyamines [11]. Linear PEI contains only secondary amines. The branched PEI has primary, secondary, and tertiary amine groups. It can be easily modified by address, imaging, or therapeutic groups [123,124]. PEI is used for colloidal stabilization in water [120], NA extraction, detection, and gene delivery [38,61,62,82,114,120,124,125,126]. Furthermore, PEI-MNC is the most efficient sorbent for ssDNA isolation in comparison to gold, silica, and graphene derivatives [114]. Due to the high positive charge in an aqueous solution, PEI interacts with the negatively charged NA and covers it. It is suitable not only for NA transport but protection from extracellular degradation by hindering its interaction with NAses [125]. An MNC coated by PEI is one of the best candidates for magnetofection, magnetic transfection, or technology [61,114]. PEI-coated MNCs could be lyophilized and are stable for not less than 4 months at room temperature, which enables the mass production of nanocomposites for various applications.
Surface functional groups provide an attractive ionic, hydrogen bond, or hydrophobic interaction with the desired target. In addition to the physical sorption, the chemical modification makes any covalent binding with the reporter, fluorescence, or “pull-out” groups probable (Figure 3) [25,81,127,128]. One of the most popular is the acylation of amino and hydroxyl group-containing polymers forming stable amide and ester bonds. However, the UV-induced immobilization of the oligonucleotides on the polymers with the primary amino groups, such as nylon-6, is possible [19,129]. The presence of ten thymidine fragments in the oligonucleotide structure allows for the easy covalent capture on the nylon-6 surface with a good yield. Such an approach was used to target NA capture from the mixture with a high specificity [19]. The possibility of selective NA isolation provides rich information about the organism in health and disease. The polymer chemical modification by “pull-out” groups represents numerous biomedical applications of NA systems [19]. For example, aptamers for various targets can be non-covalently adsorbed or chemically bound with the polymer coating. Aptamers are short sequences of NA that bind a specific target molecule. Such artificial NA is widely used in biological research laboratories and medical tests. They can show strong binding to their target, which is the reason for their name “chemical antibodies”. Aptamers can be used to identify various disease markers, be part of the drug delivery system, or have a therapeutic application [130].
Polymer-based MNCs may have various structures, which are presented in Figure 4. Core–shell, magnetic fibers, self-assembly, and doping MNCs are the most popular. Core–shell types are used everywhere, from biosensors to the theranostic area [16,44,47,131,132,133,134,135]. Polymer-based MNC nanofibers can be readily customized to adapt to different applications. It may be a small-length material with a magnetic inclusion, which may be useful for NA capture procedures due to its easy precipitation and good magnetization [19]. Furthermore, the fibers can be long, with a good porous structure and high surface area for drug absorption. Such MNCs demonstrate potential for cell and tissue adhesion and drug-loading for the formation of electrospinning biocompatible material [113]. Despite the procedure’s high costs, expensive equipment, and laborious nature, it has an excellent potential for tissue engineering approaches. The sufficient magnetization of MNCs is essential for the formation of the dimensional fiber structure. Moreover, a controlled drug release is possible via an external magnetic field through the hyperthermia effect.

2.2. Biomolecule and Biopolymer-Coated MNCs (Bioinspired MNCs)

Nanocarriers provide new possibilities in NA and gene delivery. Recently, novel strategies such as bioinspired surface coating, coating functionalization with address molecules, cell-penetrating peptides, and reporter groups have emerged [44,136]. Biomolecules and biopolymers provide selective delivery, specific binding, high biocompatibility, biodegradation, no immune response, etc. Furthermore, biopolymers have amino, carboxyl, and hydroxyl groups, which can be successfully used for additional modification toward materials with better properties and performance than the initial. Despite the advantages, MNCs are complex and may not be adapted for mass production, have a high price, low synthesis yield, and structure reproducibility. Moreover, bioinspired MNCs are not required for any bioapplication related to NA. For example, NA is easily isolated by polymer- or silica-coated MNCs. We believe that bioinspired material has more possibilities for the human organism-related areas. Among these possibilities are NA delivery, tissue engineering, in vivo diagnostics, therapy, and theranostics.
Polysaccharides are very popular for MNC surface modification [5,70,120,136,137]. The most common polysaccharides are chitosan, hyaluronic acid, heparin, starch, cellulose, agarose, and dextran. Chitosan is one of the most widely used biopolymers for MNP core stabilization [137]. It is a non-toxic, biocompatible, and biodegradable biopolymer with proven antiviral, anti-inflammatory, and antibacterial activity [11,137]. Free amine groups of chitosan allow for easy acylation, yielding the material with fluorescence, reporter, or other type groups. Due to the positive charge of amino groups, chitosan can interact with negatively charged NA and cell membranes [137,138]. ChitosanMNCs are the most widely used polysaccharides for MRI, gene delivery, NA extraction, and drug/NA systems for therapy [137,138]. Other polysaccharides, such as alginate, dextran, hyaluronic acid, heparin, mannan, pullulan, and starch, have neutral or negative charges (-OH and COOH groups), which are not suitable for efficient NA interaction. Surface chemical modification or positive polymer coating is required for magnetofection [130,137].
Protein coating provides biocompatibility, biodegradability, and less immunogenicity of MNCs [23,25,139,140,141,142,143]. For example, human serum albumin (HSA) [144,145,146] coating increases the colloidal stability in an aqueous solution, prolongs blood circulation time, prevents aggregation, and non-specific adsorption of blood components [23,25,143,147,148,149,150,151]. As one of the major plasma proteins, albumin corona forms on all nanoparticles in the bloodstream [152]. HSA increases the efficiency of tissue and cell targeting [153,154,155]. Albumin interaction with various receptors provides targeted delivery to tumors [25,143]. Moreover, HSA may be chemically modified by vitamins [156], cell-penetrating peptides, antibodies [157,158], imaging probes [159,160,161,162], NA, and drugs [156,161] for enhanced tumor delivery and theranostics [25]. AlbuminMNCs have been employed to deliver plasmids, oligonucleotides, and siRNAs for therapeutic applications [163,164].
Polydopamine (PDA) is a very important polymer that can be easily coated onto MNPs to form a uniform core–shell nanostructure [28,165]. PDA is the final oxidation product of dopamine or other catecholamines, attracting much attention as a versatile coating for MNPs [27,28,165,166,167,168,169,170,171]. PDA shows structural flexibility and strong adhesion to all types of substrates due to various functional groups. Moreover, PDA’s chemical groups allow for decoration with biomolecules and various reporter residues, achieving hybrid smart systems. Various PDA-MNCs have been reported, and their applications in bioareas have been discussed [28]. PDA is an ideal candidate for environmental remediation (the removal of pollutants) [27,28], biomedicine and imaging [28], drug and gene delivery [28,168], DNA extraction and detection [169,171], and cell tracking [167].
In recent years, there has been a growing interest in the synthesis of MNCs with biomolecules for various applications [136]. The binding of proteins, polysaccharides, lipids, peptides, and nucleic acids onto MNCs provides a bioinspired interaction with a living organism. Among the nanocomposites, biopolymer-coated MNCs are one of the most seen in various biomedical applications [25,40,102,128,136].

2.3. Inorganic Compound-Coated MNPs (Noble Metals, Silica, Calcium Carbonate, Carbon, etc.)

One of the main categories of MNCs is inorganic material coated. Multiple structures, including core–shell, Janus, and dumbbell-shaped, may be formed using different types of inorganic compounds. We present the primary coating as a silica and noble metal (usually gold) and some rare calcium carbonate, carbon, and metal–organic frameworks regarding NA-based applications.
Silica coating on MNPs is the most popular due to its low cost, stability in aqueous solution, and biocompatibility [66,102,125,172,173,174,175]. Silica shells are primarily synthesized through simple sol–gel reactions. The most common silicate precursor used in silica coating is tetraethyl orthosilicate (TEOS) (Figure 5). For the reaction, MNPs, ammonia or NaOH, and TEOS are used. Many factors, such as surfactant, solvents, temperature, and stirring speed, influence the silicaMNCs’ type, size, shape, and porosity [176]. Silica coating highly increases the stability of the magnetic core and reduces the interaction with oxygen and the formation of the reactive oxygen species’ toxicity [16,102,177,178]. Due to the porous structure, MNCs are used for drug delivery and NA capture without modification [62,66,102,114,125,176]. However, MNCs can be easily functionalized by a standard procedure, forming amino groups on the surface (Figure 5). As mentioned above, the positive charge of amino groups may be useful for NA capture [114]. A further amino group modification greatly extends the applicability of MNCs [179].
Silica surfaces may be coated by polymers, drug-loaded, and modified with various address and reporter groups [125,175,179]. The preparation of hybrid drugs and NA silica nanoparticles is a well-known strategy for enhancing the efficiency of the treatment [125,180]. Combination therapies of folate-functionalized silicaMNCs loaded with VEGF shRNA and doxorubicin show a high potential for cancer treatment [18,125]. SilicaMNCs are a widely known technology for the synthesis of hybrid multifunctional smart materials [125,175,179,180]. SilicaMNCs coated with avidin and streptavidin proteins are commonly known as “magnetic beads” for specific NA capture [82,181]. The artificial biotinylated oligonucleotide forms a duplex with a targeted NA mixture. Afterward, the protein on the silicaMNCs’ surface binds the biotin residue, forming a high-affinity complex. The specific NA may be isolated by a magnetic separation procedure with subsequent washing. Such avidin- or streptavidin-coated MNCs are commercially available elsewhere.
Noble metals are highly chemically inert materials used for surface protection. Related to MNPs, a noble metal such as gold protects the magnetic core against oxidation, corrosion, and aggregation and increases its biocompatibility [12,13,102,182,183,184,185,186]. The silver coating provides well-known antibacterial activities [186,187,188]. Indeed, the combination of MNPs and gold is suitable for hyperthermia, radionuclide (198Au) and chemotherapeutic drug cancer treatment, tumor imaging by MRI, targeted delivery of NA and drugs, and NA detection [13,70,182,183,184,185,189,190]. The synthesis of noble metal composites can be divided into chemical and physical (Figure 6). Chemical methods are used for the gold deposition on MNPs, resulting in hybrids with chemically inert surfaces. These materials are more suitable for biomedical applications than those obtained by physical approaches. The physical method of the fabrication of noble metal MNPs involves the use of laser pulse energy to transform noble metals from macrostructures into small-sized powder.
The optical properties with the strong adsorptive ability of a gold coating enable the utilization of AuMNCs for NA detection and delivery. These unique properties provide the fluorometric discrimination of mismatch DNA [191,192], PCR assay for CYP2C19 genotyping [193], methylated DNA detection by PCR [194], and single methyl discrimination in DNA aptamers [189]. AuMNCs may be used for NA magnetic separation [70,114,195]. However, without additional coating, AuMNCs have a relatively low capacity in comparison to PEIMNCs or silicaMNCs [114]. Recently, some research groups have reported the synthesis of polymer-coated AuMNCs, including polymers such as PEI, poly(acrylic acid), poly-l-lysine, or dextran [13,14,190,196]. Polymers may be used as an intermediate layer to improve the stability of the core for the fabrication of multilayer MNCs. The external polymer coating, such as PEI, can reduce and stabilize the Au shell [14]. Moreover, PEI’s positive surface charge is suitable for the intracellular delivery of siRNA or drugs (doxorubicine) [14].
Carbon and its derivatives (Figure 7) are used for the synthesis of MNCs [43,83,197,198]. Carbon species demonstrate high intrinsic electrical conductivity, excellent stability, and extensive NA adsorption suitable for biosensor production [43,83,102,199,200,201]. Carbon-coated MNPs have a relatively high magnetic moment in comparison to other forms. However, carbonMNCs usually have extremely low solubility in water and colloidal stability. Furthermore, the size and shape of the nanoparticles are heterogeneous, and the synthesis control is complicated. NA interacts with the carbon surface through π stacking, as shown for carbon nanotubes, graphene oxide, and C60 fullerene [62,199,202]. However, the combination of polymer coating and carbonMNCs leads to solubility in water and stable systems for targeted drug/gene delivery [120,198,202,203].
Calcium carbonate (CaCO3) is a well-known mineral and a highly biocompatible material. Because of the porous structure, pH-sensitivity, and biodegradability, CaCO3 nanoparticles are developed as the nanoplatform for various forms of compound isolation and drug delivery [204,205,206,207,208,209]. CaCO3 is highly sensitive to acidic pH, facilitating drug release in tumor media [204,205,206]. CaCO3 nanoparticles can bind with NA, making NA isolation and gene delivery possible [205]. For positively charged siRNA-loaded CaCO3 nanoparticles, the significantly decreased proliferation of tumor cells was shown [204,205,210,211]. Recently, Fe3O4@CaCO3 nanocomposites (CaCO3MNCs) were developed [21,84,212,213,214,215,216,217,218]. CaCO3MNCs show a high potential for ions, dyes, drug adsorption, and theranostic application [20,21,84,211,212,213]. PEI-coated or positively charged surfactant CaCO3MNCs provide enzyme immobilization [214], negatively charged microalgae [215], and cell adsorption [21]. Coating by positive-charged CaCO3MNCs may be a possibility for NA magnetic separation and gene delivery approaches.
Metal–organic frameworks (MOFs) are hybrid materials that consist of metal ions or clusters coordinated to organic ligands forming three-dimensional structures. Recently, magnetic MOF nanocomposites (MOF-MNCs) were developed for a wide range of applications such as drug and gene delivery, NA, proteins, and other biomolecule sensors, and magnetic separation [102,203,219,220,221,222]. The combination of MOF-MNCs with aptamers or other types of NA can be utilized for the fabrication of specific biosensors and theranostic systems [203,219].

3. Biomedical Applications of Magnetic Nanocomposites

MNCs have been used for various analytical applications (biosensors, magnetic separation), drug and gene delivery, imaging, and theranostics (Figure 8). NA with MNCs display unique and bioinspired properties due to the synergistic effect. As shown before, the coating of MNCs may be inorganic, organic, bioinspired, or complex. For each coating, a specific application is possible. This section presents the primary applications of NA-based MNCs. The specific information was discussed in Section 2 concerning the type of MNCs and their application.

3.1. Toxicity of MNCs

The toxicity of MNPs is an essential factor for future healthcare applications [31,223,224]. However, the studies on MNCs’ creation and safety assessment remained largely divided. While most of the safety studies have been focused on easy nanoparticles on the cell model, the material area goes forward with various smart constructions. However, most works usually present a primary cytotoxicity test, using an MTT compound on cancer cell lines [225,226]. The MTT assay does not show the interaction with blood proteins, tissue media, and delayed toxicity of the degraded product. The cancer cells are highly adapted to ROS levels and unfavorable media conditions and have activated cell growth and survival systems. In this way, the toxicity of the MNCs is still less explored [223,224,227,228,229]. Recently, studies on the toxicity of MNCs on spheroids were highlighted [185,230,231,232]. Such three-dimensional (3D) cell aggregates can mimic the tumor microenvironment. In recent years, significant progress in the development of spheroids for use as a tumor model has been obtained. The cell viability in two-dimensional (2D) cell culture monolayers and spheroids has not shown the same results [230,232,233,234]. Therefore, dose predictions from conventional cell experiments are often misleading for in vivo applications. Spheroids are a successful replacement for expensive and unethical animal experiments. However, extensive further studies are required for the stable manufacturing of various cancer spheroids and better tumor mimicking as sustainable cell growth, proliferating and non-proliferating cells, a hypoxic center, etc. [234,235].
Herein, the possible cytotoxicity and organ-specific toxicity are discussed. A high indestructibility in biological liquids or low toxicity is not required for some purposes. For example, for magnetic separation, MNCs usually do not interact with human organisms and should only have good magnetic properties and the ability to interact with a target. For in vivo studies, MNCs must be non-toxic, stable in biological liquids, and biocompatible. Unstable MNPs may be extremely toxic due to the formation of reactive oxygen species (ROS), injury to the immune system, metabolic disorders, decrease in growth rate, or changes in alterations, inflammation, ulceration, etc. (Figure 9) [44,223,224,236]. A high ROS level leads to mitochondrial membrane damage, harmful cell proliferation, modulating gene transcription, dysregulation of ion channels, RNA destruction, and DNA and lipids oxidation with a subsequent formation of a point mutation. In extreme cases, significant damage leads to cell death (Figure 9) [228,236]. The oxidative stress mechanism can be provided by the release of ferrous ions due to the instability of the MNCs, direct ROS generation on the MNCs’ surface, and altering the mitochondrial function and signaling pathways [228,236].
The MNCs’ size, shape, surface charge, coating, and surface modification highly influence biodistribution and toxicity [223]. MNCs with a size less than 10 nm are quickly removed through renal clearance. Nanocomposites greater than 180–200 nm are filtrated by the spleen. Therefore, MNCs in the range from 10 to 150 nm are the most preferred. The surfaces of nanoparticles are rapidly covered in blood by various proteins. Surface chemistry highly influences nanocomposite biodistribution. For example, albumin coating usually prevents non-specific blood interaction and liver accumulation, prolongs circulation time, and moderates particle uptake in cancer tissue [25]. The wrong coating of the nanocomposites may lead to the MNCs’ destabilization, aggregation, and precipitation [29]. Inside the body, MNCs can be absorbed through interactions with proteins (e.g., protein albumin), blood components, and cells [228]. Blood compatibility is essential for any in vivo application of MNCs. Lack of stability in the blood can trigger liver accumulation with further degradation and elimination from the body. In the worst case, the co-coagulation and precipitation of MNCs and the blood component may happen, activating thrombus formation.
The MNCs’ dose, initial concentration, biodistribution, and circulation time should be taken into account [42]. The nanocomposites can be distributed into various organs and further metabolized. The overload of local MNCs may lead to high levels of free iron ion release and ROS generation in the tissue and cause aberrant cellular reactions and organ-specific toxicity [223,224,237]. The in vivo toxicity experiment is an expensive and vast work, which has greatly shut down the progress in this area [224,228]. Almost all organs are influenced by the toxic effects of MNCs. Among those are the heart, lungs, liver, kidney, and nervous and reproductive systems. For example, heart-specific toxicity provides contractile apparatus and endothelial damage, the violation of a conducting system, and ischemia [224]. The MNCs’ organ-specific toxic effects may be associated with ROS generation, leading to changes in glutathione, superoxide dismutase, and coenzyme levels [224]. Afterward, the in vivo interaction of MNCs and the biological system is quite complicated and dynamic. In the future, extended toxicity studies could help to bridge the gap between in vitro research results and successful clinical trials.

3.2. Drug and Gene Delivery, Therapy, and Diagnostics (Theranostics)

Nanoparticles and MNCs became extremely popular for cancer therapy due to the possible targeted delivery. MNCs provide drug or gene delivery to the cell or tissue by an external magnetic field. Magnetic transfection, or magnetofection, is a method that uses magnetic fields to transport NA-based MNCs to target cells. Magnetofection has been adapted to various NA types, including aptamers, siRNA, miRNA, shRNA, etc. [76,77,78,238,239,240]. Such technology is a possibility to solve the drug resistance problem and the low efficiency of gene delivery through cell membranes [241]. A combination of MNCs and siRNA or antisense oligonucleotides may be successfully used instead of a chemotherapeutic drug, resulting in a therapeutic effect [76,238,242,243,244,245].
As stated above, the magnetic core of MNCs has multimodal advantages, such as possible tracking by magnetic resonance imaging (MRI) or magnetic particle imaging (MPI) and the hyperthermia effect. MRI and MPI are great non-invasive diagnostic techniques [34,43,44,148,246,247,248,249]. MRI provides a high-resolution and easy image contrast manipulation. MNPs are usually known as T2-contrast agents, which lead to a dark zone on the MRI image. However, the previous simple MNPs were withdrawn due to the side effects. There are many successful in vitro experiments and undergoing pre-clinical animal studies for MNCs [34]. MPI is a relatively novel technology that was presented in 2005. MPI detects tracer MNCs selectively, providing the signal is observed without background with a high signal-to-noise ratio [44]. The method possesses potential for tumor, metastases, and cell detection.
The hyperthermia effect is generated by an alternating magnetic field [250,251]. In the presence of MNCs, heat appears in local regions, which damages tumor cells. The method is limited by the MNCs’ quality, size, morphology, and coating. The simultaneous drug, therapeutic NA, and hyperthermia using single-MNCs is a promising anticancer strategy. Recently, various MNCs to be used for multimodal imaging and theranostics have been developed [34,40,44]. The combination of MRI, MPI, and primary used methods, such as single-photon emission computed tomography (SPECT), computed tomography (CT), positron emission tomography (PET), and optical imaging, have become known [34,40,252,253]. However, theranostic synthesis is a complicated problem, which limits progress in the area [88].

3.3. Magnetic Separation and Biosensors

Solid-phase magnetic separation is a much more efficient protocol for NA isolation than traditional approaches [62,69,71,187,254,255,256,257,258,259,260]. Magnetic separation is a relatively cheap, quick method for pure NA capture with a high yield. It usually requires ~10 min for NA isolation from the mixture using specific MNCs and a rack with magnets (Figure 10). However, the high cost of commercially available MNCs limits their routine applications. Moreover, the specific “bind-release” of NA MNCs is extremely rare. One of the primarily used magnetic beads has an avidin/streptavidin coating and forms a specific complex with biotin-labeled oligonucleotide [261]. Biotinylated oligonucleotide interacts with a targeted NA. Such systems are specific. However, the high binding constant of avidin to biotin hinders the separation of NA from MNPs and their further use. Other commercially available MNCs for NA isolation are not specific and usually bind all the NA in the probe with a certain degree of purity. Such MNCs work on the ionic interaction of NA to the MNC’s surface. Therefore, protein binding is possible. Finally, new cheap, high-capacity sorbents for the capture of NA are required.
Recently, magnetic cell separation has become a comprehensive technology for targeted cell population separation for various applications [50]. Some MNCs for cell isolation work by using NA-based MNCs bearing aptamers. However, aptamer-based magnetic cell separation faces many obstacles, which makes it difficult to use such an approach in common practice [50]. DNA aptamers can also provide specific recognition capabilities against many targets used for magnetic separation.
Selectively sensing a single NA allows for the discovery of rich information regarding human health [71]. The separation of NA, with further analysis, is a laborious procedure that, in some cases, does not make sense. The new specific, sensitive methods are required for rapid diagnosis [49]. The surface of the specific MNCs for NA detection may be modified by NA-specific molecules, including antibodies, aptamers, NA, proteins, etc. (Figure 11). The transducer from “chemical” to “physical” signal may be electrochemical, optical, piezoelectric, etc. [49,262,263] (Figure 12). In some cases, the isolation of NA with subsequent analysis is required. This procedure may be performed using MNCs with further PCR analysis [261]. For instance, MNCs can improve the sensitivity of the PCR with an extreme detection limit. Compared to traditional PCR approaches, an MNC-based PCR shows richness and a high potential [261].

3.4. Magnetic Molecularly Imprinted Polymers

MNPs have been extensively developed for their excellent separation and extraction ability. A new innovative approach in this area is the use of MNPs with a molecularly imprinted polymer [264,265,266]. Molecular imprinting is a technique to create molecule (template)-specific cavities in polymer matrices (Figure 12).
The procedure is similar to the enzymes’ “lock and key” model. The resulting polymer is called molecularly imprinted polymers (MIPs) [266,267]. For the magnetic core, magnetic molecularly imprinted polymers (MMIPs) have been developed [264,268]. In the structure of MIPs, some regions can specifically interact with template molecules or structurally related molecules. Recognition can occur concerning the shape, size, or due to interactions between the functionalities of the template and polymer [267]. Two types of MIPs are known [269]. Among those, the classic version involves the polymerization of functional monomers, templates, and cross-linking agents (Figure 12, top line). The second one is the change in the polymer state from liquid to solid in the presence of a template.
MIPs have many advantages, such as chemical and physical stability, easy synthesis, reusability, and cost-efficient preparation [91,266]. Polymer matrices may consist of organic or inorganic compounds capable of recognizing molecules or ions [264,266,267,268]. The synthesis methods of MIPs can be classified according to the format (two- or three-dimensional) and bonding type between the functional monomers and template (covalent or non-covalent) and by the nature of the functional monomers (organic or inorganic). MIPs are used in many areas, such as synthetic recognition elements, solid-phase extraction, liquid chromatography, electrochromatography, assays, drug delivery, theranostics, and biosensor production [266,267,270,271]. MIPs are primarily prepared by bulk polymerization as monoliths (3D imprinting). The first templates for MIP recognition were low molecular weight biologically active compounds such as vitamins, hormones, toxins, drugs, nucleotides, NA, and their derivatives [272,273,274,275]. However, this method has various drawbacks, such as a low amount of binding sites near the surface, inaccessible recognition sites within the polymer bulk, a wide range of particle sizes, and non-uniform morphology [276,277]. The transition to the imprinting of biomolecules (nucleic acids, peptides, and proteins) requires significant changes in the existing imprinting protocols and the emergence of new ones [266]. Obtaining MIPs for biomolecules remains a formidable challenge due to their large dimensions, low solubility and stability, complex structure, slow mass transfer, and structural flexibility in solution. The bulk polymerization is limited to the macromolecule and biomolecule imprinting, including peptides, proteins, NA, viruses, and bacteria. Nearly 1200 research articles were published annually on MIP-based biosensors, out of which only nearly 10% included the recognition of biomacromolecules [278]. Surface molecular imprinting seems to be an alternative approach that can address some of the shortcomings of the synthesis of a primary MIP. Inorganic materials, such as silica, magnetic, gold, and silver nanoparticles, are especially widely used as a core for MIPs [279,280]. The combination of MIPs and other materials combines features yielding smart core–shell MIP structures which allow for the control of the size and distribution of the synthesis. The hybrid magnetic MIP (MMIP) has the advantages of the technology of MIPs and MNPs.
The magnetic properties of the MMIP allow for magnetic separation, imaging, hyperthermia, and selective template release [264,265,268,281,282,283]. MMIPs have shown a high potential in identifying a broad spectrum of analytes, from small molecule enantiomers to large proteins, NA, and macromolecules [91,264,284,285]. The possibility to automatize this process by using magnetic properties is also an interesting feature for industrialization and mass production. Nowadays, MMIPs are widely used in various fields of biomedicine, such as biosensors, drug and gene delivery, and NA isolation [267,279,285]. MMIPs show a high potential for use in cancer therapy due to targeted delivery by an external magnetic field, hyperthermia effect, and possible simultaneous drug and therapeutic NA delivery [286]. The use of MMIPs for NA-based applications is also being extensively studied [91,267,287,288,289]. NA can act as both templates and complex macromolecular functional monomers, which provide unique properties to the resulting MIPs [290]. Consequently, combining MIPs with NA with magnetic properties provides a new class of smart synthetic NA receptors, i.e., NA-MMIPs [287]. These materials open up new possibilities in this research area. Table 1 summarizes NA-based MNC biomedical applications.

4. Conclusions and Future Prospects

The magnetic core on MNCs is a promising feature for unique properties and various applications. It endows the possibility of MRI and MPI diagnostics, magnetic separation, hyperthermia, and targeting by an external field drug/gene delivery. Combining with other types of coating is vital to develop MNCs with diversified properties. Finally, MNCs are a promising core for the creation of a new smart construction. The possible surface functionalization opens up numerous tool options, which could greatly advance the field. We believe that bioinspired MNCs have a great proven potential for cancer treatment and biosensing applications. NA-based MNCs take an important place in analytical applications, biosensing, magnetic separation, diagnostics, therapy, and theranostic areas. Recently, next-generation constructions were developed and combined with one nano platform therapeutic NA, drug, hyperthermia possibilities, imaging, and targeted delivery. Such incredible progress makes for a bright future in cancer treatment. New rapid and efficient synthesis procedures for MNCs are required for future studies. Despite the widespread advances of MNCs, nanomedicine and toxicology integration proved to be essential steps toward in vivo applications. In this regard, a collaboration of chemists, physicists, biologists, and physicians is extremely necessary for further breakthrough developments. However, new simple, cheap, and efficient toxicity models and clinical protocols need to be clearly defined. This gap in the toxicity evaluation is a great limiting factor for practical usage. Finally, NA-based MNCs will rightfully take their place among various biomedical applications. We certainly hope that the improvement in the synthesis and functions of bioinspired MNCs will afterward lead to a new era of nanomedicine.

Author Contributions

Conceptualization, E.D. and A.C.; writing—original draft preparation, E.D., V.P. and A.C.; project administration, A.C.; funding acquisition, A.C. and E.D. All authors have read and agreed to the published version of the manuscript.

Funding

This study was funded by the Russian Science Foundation (grant no. 21-64-00017) and the polymer-coated nanocomposites partially by the Ministry of Science and Higher Education of the Russian Federation (state registration No. 121031300042-1).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Das, A.K.; Fanan, A.; Ali, D.; Solanki, V.S.; Pare, B.; Almutairi, B.O.; Agrawal, N.; Yadav, N.; Pareek, V.; Yadav, V.K. Green Synthesis of Unsaturated Fatty Acid Mediated Magnetite Nanoparticles and Their Structural and Magnetic Studies. Magnetochemistry 2022, 8, 174. [Google Scholar] [CrossRef]
  2. Zaloga, J.; Feoktystov, A.; Garamus, V.M.; Karawacka, W.; Ioffe, A.; Brückel, T.; Tietze, R.; Alexiou, C.; Lyer, S. Studies on the Adsorption and Desorption of Mitoxantrone to Lauric Acid/Albumin Coated Iron Oxide Nanoparticles. Colloids Surf. B Biointerfaces 2018, 161, 18–26. [Google Scholar] [CrossRef] [PubMed]
  3. Shete, P.B.; Patil, R.M.; Tiwale, B.M.; Pawar, S.H. Water dispersible oleic acid-coated Fe3O4 nanoparticles for biomedical applications. J. Magn. Magn. Mater. 2015, 377, 406–410. [Google Scholar] [CrossRef]
  4. Mahdavi, M.; Ahmad, M.B.; Haron, M.J.; Namvar, F.; Nadi, B.; Ab Rahman, M.Z.; Amin, J. Synthesis, surface modification and characterisation of biocompatible magnetic iron oxide nanoparticles for biomedical applications. Molecules 2013, 18, 7533–7548. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Darwish, M.S.A.; Mostafa, M.H.; Al-Harbi, L.M. Polymeric Nanocomposites for Environmental and Industrial Applications. Int. J. Mol. Sci. 2022, 23, 1023. [Google Scholar] [CrossRef]
  6. Mukhopadhyay, A.; Joshi, N.; Chattopadhyay, K.; De, G. A facile synthesis of PEG-coated magnetite (Fe3O4) nanoparticles and their prevention of the reduction of cytochrome C. ACS Appl. Mater. Interfaces 2012, 4, 142–149. [Google Scholar] [CrossRef]
  7. Huang, Y.; Zhang, B.; Xie, S.; Yang, B.; Xu, Q.; Tan, J. Superparamagnetic Iron Oxide Nanoparticles Modified with Tween 80 Pass through the Intact Blood-Brain Barrier in Rats under Magnetic Field. ACS Appl. Mater. Interfaces 2016, 8, 11336–11341. [Google Scholar] [CrossRef]
  8. Yoon, H.M.; Kang, M.S.; Choi, G.E.; Kim, Y.J.; Bae, C.H.; Yu, Y.B.; Jeong, Y. Il Stimuli-responsive drug delivery of doxorubicin using magnetic nanoparticle conjugated poly(Ethylene glycol)-g-chitosan copolymer. Int. J. Mol. Sci. 2021, 22, 13169. [Google Scholar] [CrossRef]
  9. Snoderly, H.T.; Freshwater, K.A.; Martinez de la Torre, C.; Panchal, D.M.; Vito, J.N.; Bennewitz, M.F. PEGylation of Metal Oxide Nanoparticles Modulates Neutrophil Extracellular Trap Formation. Biosensors 2022, 12, 123. [Google Scholar] [CrossRef]
  10. Kovrigina, E.; Chubarov, A.; Dmitrienko, E. High Drug Capacity Doxorubicin-Loaded Iron Oxide Nanocomposites for Cancer Therapy. Magnetochemistry 2022, 8, 54. [Google Scholar] [CrossRef]
  11. Mylkie, K.; Nowak, P.; Rybczynski, P.; Ziegler-Borowska, M. Polymer-coated magnetite nanoparticles for protein immobilization. Materials 2021, 14, 248. [Google Scholar] [CrossRef]
  12. Schwaminger, S.P.; Bauer, D.; Fraga-García, P. Gold-iron oxide nanohybrids: Insights into colloidal stability and surface-enhanced Raman detection. Nanoscale Adv. 2021, 3, 6438–6445. [Google Scholar] [CrossRef]
  13. Tarkistani, M.A.M.; Komalla, V.; Kayser, V. Recent advances in the use of iron–gold hybrid nanoparticles for biomedical applications. Nanomaterials 2021, 11, 1227. [Google Scholar] [CrossRef]
  14. Elmi, G.R.; Saleem, K.; Baig, M.M.F.A.; Aamir, M.N.; Wang, M.; Gao, X.; Abbas, M.; Rehman, M.U. Recent Advances of Magnetic Gold Hybrids and Nanocomposites, and Their Potential Biological Applications. Magnetochemistry 2022, 8, 38. [Google Scholar] [CrossRef]
  15. Buema, G.; Herea, D.; Chiriac, H.; Lupu, N.; Minuti, A.E.; Stoian, G.; Shore, D.; Pierre, V.C.; Tabakovic, I.; Stadler, B.J.H. Synthesis and Characterization of Gold-Shell Magnetic Nanowires for Theranostic Applications. Coatings 2022, 12, 1755. [Google Scholar]
  16. Zelepukin, I.V.; Shipunova, V.O.; Mirkasymov, A.B.; Nikitin, P.I.; Nikitin, M.P. Synthesis and Characterization of Hybrid Core-Shell Fe3O4/SiO2 Nanoparticles for Biomedical Applications. Acta Nat. 2017, 9, 58–65. [Google Scholar] [CrossRef] [Green Version]
  17. Turrina, C.; Oppelt, A.; Mitzkus, M.; Berensmeier, S.; Schwaminger, S.P. Silica-coated superparamagnetic iron oxide nanoparticles: New insights into the influence of coating thickness on the particle properties and lasioglossin binding. MRS Commun. 2022, 12, 632–639. [Google Scholar] [CrossRef]
  18. Li, T.; Shen, X.; Geng, Y.; Chen, Z.; Li, L.; Li, S.; Yang, H.; Wu, C.; Zeng, H.; Liu, Y. Folate-Functionalized Magnetic-Mesoporous Silica Nanoparticles for Drug/Gene Codelivery to Potentiate the Antitumor Efficacy. ACS Appl. Mater. Interfaces 2016, 8, 13748–13758. [Google Scholar] [CrossRef]
  19. Bulgakova, A.; Chubarov, A.; Dmitrienko, E. Magnetic Nylon 6 Nanocomposites for the Microextraction of Nucleic Acids from Biological Samples. Magnetochemistry 2022, 8, 85. [Google Scholar] [CrossRef]
  20. Vavaev, E.S.; Novoselova, M.; Shchelkunov, N.M.; German, S.; Aleksei, S.; Mokrousov, M.D.; Zelepukin, I.V.; Burov, A.M.; Khlebtsov, B.N.; Lyubin, E.V.; et al. CaCO3 Nanoparticles Coated with Alternating Layers of Poly-L-Arginine Hydrochloride and Fe3O4 Nanoparticles as Navigable Drug Carriers and Hyperthermia Agents. ACS Appl. Nano Mater. 2022, 5, 2994–3006. [Google Scholar] [CrossRef]
  21. Wang, P.; Xue, J.; Wu, S.; Pei, Y.; Xu, L.; Wang, Y. Cell-Friendly Isolation and pH-Sensitive Controllable Release of Circulating Tumor Cells by Fe3O4@CaCO3 Nanoplatform. Adv. Mater. Interfaces 2021, 8, 2101191. [Google Scholar] [CrossRef]
  22. Schubert, J.; Chanana, M. Coating Matters: Review on Colloidal Stability of Nanoparticles with Biocompatible Coatings in Biological Media, Living Cells and Organisms. Curr. Med. Chem. 2018, 25, 4553–4586. [Google Scholar] [CrossRef] [PubMed]
  23. Khramtsov, P.; Barkina, I.; Kropaneva, M.; Bochkova, M.; Timganova, V.; Nechaev, A.; Byzov, I.; Zamorina, S.; Yermakov, A.; Rayev, M. Magnetic nanoclusters coated with albumin, casein, and gelatin: Size tuning, relaxivity, stability, protein corona, and application in nuclear magnetic resonance immunoassay. Nanomaterials 2019, 9, 1345. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Schwaminger, S.P.; Blank-Shim, S.A.; Scheifele, I.; Pipich, V.; Fraga-García, P.; Berensmeier, S. Design of Interactions Between Nanomaterials and Proteins: A Highly Affine Peptide Tag to Bare Iron Oxide Nanoparticles for Magnetic Protein Separation. Biotechnol. J. 2019, 14, 1800055. [Google Scholar] [CrossRef]
  25. Chubarov, A.S. Serum Albumin for Magnetic Nanoparticles Coating. Magnetochemistry 2022, 8, 13. [Google Scholar] [CrossRef]
  26. Vismara, E.; Bongio, C.; Coletti, A.; Edelman, R.; Serafini, A.; Mauri, M.; Simonutti, R.; Bertini, S.; Urso, E.; Assaraf, Y.G.; et al. Albumin and hyaluronic acid-coated superparamagnetic iron oxide nanoparticles loaded with paclitaxel for biomedical applications. Molecules 2017, 22, 1030. [Google Scholar] [CrossRef] [Green Version]
  27. Li, Y.; Huang, L.; He, W.; Chen, Y.; Lou, B. Preparation of functionalized magnetic Fe3O4@Au@polydopamine nanocomposites and their application for copper(II) removal. Polymers 2018, 10, 570. [Google Scholar] [CrossRef] [Green Version]
  28. Siciliano, G.; Monteduro, A.G.; Turco, A.; Primiceri, E.; Rizzato, S.; Depalo, N.; Curri, M.L.; Maruccio, G. Polydopamine-Coated Magnetic Iron Oxide Nanoparticles: From Design to Applications. Nanomaterials 2022, 12, 1145. [Google Scholar] [CrossRef]
  29. Issa, B.; Obaidat, I.M.; Albiss, B.A.; Haik, Y. Magnetic nanoparticles: Surface effects and properties related to biomedicine applications. Int. J. Mol. Sci. 2013, 14, 21266–21305. [Google Scholar] [CrossRef] [Green Version]
  30. Kudr, J.; Haddad, Y.; Richtera, L.; Heger, Z.; Cernak, M.; Adam, V.; Zitka, O. Magnetic nanoparticles: From design and synthesis to real world applications. Nanomaterials 2017, 7, 243. [Google Scholar] [CrossRef]
  31. Petrov, K.D.; Chubarov, A.S. Magnetite Nanoparticles for Biomedical Applications. Encyclopedia 2022, 2, 1811–1828. [Google Scholar] [CrossRef]
  32. Anderson, S.D.; Gwenin, V.V.; Gwenin, C.D. Magnetic Functionalized Nanoparticles for Biomedical, Drug Delivery and Imaging Applications. Nanoscale Res. Lett. 2019, 14, 1–16. [Google Scholar] [CrossRef] [Green Version]
  33. Lamichhane, N.; Sharma, S.; Parul; Verma, A.K.; Roy, I.; Sen, T. Iron oxide-based magneto-optical nanocomposites for in vivo biomedical applications. Biomedicines 2021, 9, 288. [Google Scholar] [CrossRef]
  34. Creţu, B.E.B.; Dodi, G.; Shavandi, A.; Gardikiotis, I.; Şerban, I.L.; Balan, V. Imaging constructs: The rise of iron oxide nanoparticles. Molecules 2021, 26, 3437. [Google Scholar] [CrossRef]
  35. Anik, M.I.; Hossain, M.K.; Hossain, I.; Mahfuz, A.M.U.B.; Rahman, M.T.; Ahmed, I. Recent progress of magnetic nanoparticles in biomedical applications: A review. Nano Sel. 2021, 2, 1146–1186. [Google Scholar] [CrossRef]
  36. Mittal, A.; Gandhi, S.; Roy, I. Mechanistic interaction studies of synthesized ZIF-8 nanoparticles with bovine serum albumin using spectroscopic and molecular docking approaches. Sci. Rep. 2022, 12, 10331. [Google Scholar] [CrossRef]
  37. Comanescu, C. Magnetic Nanoparticles: Current Advances in Nanomedicine, Drug Delivery and MRI. Chemistry 2022, 4, 872–930. [Google Scholar] [CrossRef]
  38. Włodarczyk, A.; Gorgoń, S.; Radoń, A.; Bajdak-Rusinek, K. Magnetite Nanoparticles in Magnetic Hyperthermia and Cancer Therapies: Challenges and Perspectives. Nanomaterials 2022, 12, 1807. [Google Scholar] [CrossRef] [PubMed]
  39. Materón, E.M.; Miyazaki, C.M.; Carr, O.; Joshi, N.; Picciani, P.H.S.; Dalmaschio, C.J.; Davis, F.; Shimizu, F.M. Magnetic nanoparticles in biomedical applications: A review. Appl. Surf. Sci. Adv. 2021, 6, 100163. [Google Scholar] [CrossRef]
  40. Mittal, A.; Roy, I.; Gandhi, S. Magnetic Nanoparticles: An Overview for Biomedical Applications. Magnetochemistry 2022, 8, 107. [Google Scholar] [CrossRef]
  41. Chouhan, R.S.; Horvat, M.; Ahmed, J.; Alhokbany, N.; Alshehri, S.M.; Gandhi, S. Magnetic nanoparticles—A multifunctional potential agent for diagnosis and therapy. Cancers 2021, 13, 2213. [Google Scholar] [CrossRef]
  42. Shabatina, T.I.; Vernaya, O.I.; Shabatin, V.P.; Melnikov, M.Y. Magnetic nanoparticles for biomedical purposes: Modern trends and prospects. Magnetochemistry 2020, 6, 30. [Google Scholar] [CrossRef]
  43. Ganapathe, L.S.; Mohamed, M.A.; Yunus, R.M.; Berhanuddin, D.D. Magnetite (Fe3O4) nanoparticles in biomedical application: From synthesis to surface functionalisation. Magnetochemistry 2020, 6, 68. [Google Scholar] [CrossRef]
  44. Hepel, M. Magnetic nanoparticles for nanomedicine. Magnetochemistry 2020, 6, 3. [Google Scholar] [CrossRef] [Green Version]
  45. Dulińska-Litewka, J.; Łazarczyk, A.; Hałubiec, P.; Szafrański, O.; Karnas, K.; Karewicz, A. Superparamagnetic iron oxide nanoparticles-current and prospective medical applications. Materials 2019, 12, 617. [Google Scholar] [CrossRef] [Green Version]
  46. Stueber, D.D.; Villanova, J.; Aponte, I.; Xiao, Z. Magnetic Nanoparticles in Biology and Medicine: Past, Present, and Future Trends. Pharmaceutics 2021, 13, 943. [Google Scholar] [CrossRef]
  47. Socoliuc, V.; Peddis, D.; Petrenko, V.I.; Avdeev, M.V.; Susan-Resiga, D.; Szabó, T.; Turcu, R.; Tombácz, E.; Vékás, L. Magnetic nanoparticle systems for nanomedicine—A materials science perspective. Magnetochemistry 2020, 6, 2. [Google Scholar] [CrossRef] [Green Version]
  48. Bruschi, M.L.; de Toledo, L.d.A.S. Pharmaceutical applications of iron-oxide magnetic nanoparticles. Magnetochemistry 2019, 5, 50. [Google Scholar] [CrossRef] [Green Version]
  49. Krishnan, S.; Goud, K.Y. Magnetic Particle Bioconjugates: A Versatile Sensor Approach. Magnetochemistry 2019, 5, 64. [Google Scholar] [CrossRef] [Green Version]
  50. Frenea-Robin, M.; Marchalot, J. Basic Principles and Recent Advances in Magnetic Cell Separation. Magnetochemistry 2022, 8, 11. [Google Scholar] [CrossRef]
  51. Mariño, M.A.; Fulaz, S.; Tasic, L. Magnetic nanomaterials as biocatalyst carriers for biomass processing: Immobilization strategies, reusability, and applications. Magnetochemistry 2021, 7, 133. [Google Scholar] [CrossRef]
  52. Birringer, R.; Gleiter, H.; Klein, H.P.; Marquardt, P. Nanocrystalline materials an approach to a novel solid structure with gas-like disorder? Phys. Lett. A 1984, 102, 365–369. [Google Scholar] [CrossRef]
  53. Katz, E. Synthesis, properties and applications of magnetic nanoparticles and nanowires—A brief introduction. Magnetochemistry 2019, 5, 61. [Google Scholar] [CrossRef] [Green Version]
  54. Antone, A.J.; Sun, Z.; Bao, Y. Preparation and application of iron oxide nanoclusters. Magnetochemistry 2019, 5, 45. [Google Scholar] [CrossRef] [Green Version]
  55. Nuzhina, J.V.; Shtil, A.A.; Prilepskii, A.Y.; Vinogradov, V.V. Preclinical Evaluation and Clinical Translation of Magnetite-Based Nanomedicines. J. Drug Deliv. Sci. Technol. 2019, 54, 101282. [Google Scholar] [CrossRef]
  56. Dudchenko, N.; Pawar, S.; Perelshtein, I.; Fixler, D. Magnetite Nanoparticles: Synthesis and Applications in Optics and Nanophotonics. Materials 2022, 15, 2601. [Google Scholar] [CrossRef]
  57. Xu, S.; Lee, T.R. Fe3O4 Nanoparticles: Structures, Synthesis, Magnetic Properties, Surface Functionalization, and Emerging Applications. Appl. Sci. 2021, 11, 11301. [Google Scholar] [CrossRef]
  58. Bowers, A.N.; Trujillo-Rodríguez, M.J.; Farooq, M.Q.; Anderson, J.L. Extraction of DNA with magnetic ionic liquids using in situ dispersive liquid–liquid microextraction. Anal. Bioanal. Chem. 2019, 411, 7375–7385. [Google Scholar] [CrossRef]
  59. Vanyorek, L.; Ilosvai, Á.M.; Szőri-Dorogházi, E.; Váradi, C.; Kristály, F.; Prekob, Á.; Fiser, B.; Varga, T.; Kónya, Z.; Viskolcz, B. Synthesis of iron oxide nanoparticles for DNA purification. J. Dispers. Sci. Technol. 2021, 42, 693–700. [Google Scholar] [CrossRef]
  60. Wang, J.; Ali, Z.; Si, J.; Wang, N.; He, N.; Li, Z. Simultaneous extraction of DNA and RNA from hepatocellular carcinoma (Hep G2) based on silica-coated magnetic nanoparticles. J. Nanosci. Nanotechnol. 2017, 17, 802–806. [Google Scholar] [CrossRef]
  61. Danthanarayana, A.N.; Manatunga, D.C.; De Silva, R.M.; Chandrasekharan, N.V.; De Silva, K.M.N. Magnetofection and isolation of DNA using polyethyleneimine functionalized magnetic iron oxide nanoparticles. R. Soc. Open Sci. 2018, 5, 181369. [Google Scholar] [CrossRef]
  62. Li, P.; Li, M.; Yue, D.; Chen, H. Solid-phase extraction methods for nucleic acid separation. A review. J. Sep. Sci. 2022, 45, 172–184. [Google Scholar] [CrossRef]
  63. Tang, C.; He, Z.; Liu, H.; Xu, Y.; Huang, H.; Yang, G.; Xiao, Z.; Li, S.; Liu, H.; Deng, Y.; et al. Application of magnetic nanoparticles in nucleic acid detection. J. Nanobiotechnol. 2020, 18, 1–19. [Google Scholar] [CrossRef] [Green Version]
  64. Chacón-Torres, J.C.; Reinoso, C.; Navas-León, D.G.; Briceño, S.; González, G. Optimized and scalable synthesis of magnetic nanoparticles for RNA extraction in response to developing countries’ needs in the detection and control of SARS-CoV-2. Sci. Rep. 2020, 10, 19004. [Google Scholar] [CrossRef]
  65. Ali, T.H.; Mandal, A.M.; Heidelberg, T.; Hussen, R.S.D. Sugar based cationic magnetic core–shell silica nanoparticles for nucleic acid extraction. RSC Adv. 2022, 12, 13566–13579. [Google Scholar] [CrossRef]
  66. Bag, S.; Rauwolf, S.; Schwaminger, S.P.; Wenzel, W.; Berensmeier, S. DNA Binding to the Silica: Cooperative Adsorption in Action. Langmuir 2021, 37, 5902–5908. [Google Scholar] [CrossRef]
  67. Ma, Y.; Chen, T.; Iqbal, M.Z.; Yang, F.; Hampp, N.; Wu, A.; Luo, L. Applications of magnetic materials separation in biological nanomedicine. Electrophoresis 2019, 40, 2011–2028. [Google Scholar] [CrossRef]
  68. Marengo, A.; Cagliero, C.; Sgorbini, B.; Anderson, J.L.; Emaus, M.N.; Bicchi, C.; Bertea, C.M.; Rubiolo, P. Development of an innovative and sustainable one-step method for rapid plant DNA isolation for targeted PCR using magnetic ionic liquids. Plant Methods 2019, 15, 23. [Google Scholar] [CrossRef]
  69. Wang, L.; He, K.; Sadak, O.; Wang, X.; Wang, Q.; Xu, X. Visual detection of in vitro nucleic acid replication by submicro- and nano-sized materials. Biosens. Bioelectron. 2020, 169, 112602. [Google Scholar] [CrossRef]
  70. Sosa-Acosta, J.R.; Iriarte-Mesa, C.; Ortega, G.A.; Díaz-García, A.M. DNA–Iron Oxide Nanoparticles Conjugates: Functional Magnetic Nanoplatforms in Biomedical Applications. Top. Curr. Chem. 2020, 378, 1–29. [Google Scholar] [CrossRef]
  71. Gessner, I.; Fries, J.W.U.; Brune, V.; Mathur, S. Magnetic nanoparticle-based amplification of microRNA detection in body fluids for early disease diagnosis. J. Mater. Chem. B 2021, 9, 9–22. [Google Scholar] [CrossRef]
  72. Bobrikova, E.; Chubarov, A.; Dmitrienko, E. The Effect of pH and Buffer on Oligonucleotide Affinity for Iron Oxide Nanoparticles. Magnetochemistry 2021, 7, 128. [Google Scholar] [CrossRef]
  73. Min, J.H.; Woo, M.K.; Yoon, H.Y.; Jang, J.W.; Wu, J.H.; Lim, C.S.; Kim, Y.K. Isolation of DNA using magnetic nanoparticles coated with dimercaptosuccinic acid. Anal. Biochem. 2014, 447, 114–118. [Google Scholar] [CrossRef]
  74. Sosa-Acosta, J.R.; Silva, J.A.; Fernández-Izquierdo, L.; Díaz-Castañón, S.; Ortiz, M.; Zuaznabar-Gardona, J.C.; Díaz-García, A.M. Iron Oxide Nanoparticles (IONPs) with potential applications in plasmid DNA isolation. Colloids Surf. A Physicochem. Eng. Asp. 2018, 545, 167–178. [Google Scholar] [CrossRef]
  75. Vaughan, H.J.; Green, J.J.; Tzeng, S.Y. Cancer-Targeting Nanoparticles for Combinatorial Nucleic Acid Delivery. Adv. Mater. 2020, 32, 1901081. [Google Scholar] [CrossRef] [PubMed]
  76. Mendes, B.B.; Conniot, J.; Avital, A.; Yao, D.; Jiang, X.; Zhou, X.; Sharf-Pauker, N.; Xiao, Y.; Adir, O.; Liang, H.; et al. Nanodelivery of nucleic acids. Nat. Rev. Methods Prim. 2022, 2, 1–21. [Google Scholar] [CrossRef] [PubMed]
  77. Huang, R.-Y.; Liu, Z.-H.; Weng, W.-H.; Chang, C.-W. Magnetic nanocomplexes for gene delivery applications. J. Mater. Chem. B 2021, 9, 4267–4286. [Google Scholar] [CrossRef] [PubMed]
  78. Sizikov, A.A.; Kharlamova, M.V.; Nikitin, M.P.; Nikitin, P.I.; Kolychev, E.L. Nonviral locally injected magnetic vectors for in vivo gene delivery: A review of studies on magnetofection. Nanomaterials 2021, 11, 1078. [Google Scholar] [CrossRef] [PubMed]
  79. Prosen, L.; Prijic, S.; Music, B.; Lavrencak, J.; Cemazar, M.; Sersa, G. Magnetofection: A reproducible method for gene delivery to melanoma cells. BioMed Res. Int. 2013, 2013, 6–8. [Google Scholar] [CrossRef] [Green Version]
  80. Sizikov, A.A.; Nikitin, P.I.; Nikitin, M.P. Magnetofection In Vivo by Nanomagnetic Carriers Systemically Administered into the Bloodstream. Pharmaceutics 2021, 13, 1927. [Google Scholar] [CrossRef]
  81. Gautam, A. DNA and RNA Isolation Techniques for Non-Experts; Springer: Cham, Switzerland, 2022; ISBN 9783030942298. [Google Scholar]
  82. Berensmeier, S. Magnetic particles for the separation and purification of nucleic acids. Appl. Microbiol. Biotechnol. 2006, 73, 495–504. [Google Scholar] [CrossRef]
  83. Samanta, A.; Medintz, I.L. Nanoparticles and DNA—A powerful and growing functional combination in bionanotechnology. Nanoscale 2016, 8, 9037–9095. [Google Scholar] [CrossRef] [Green Version]
  84. Xue, J.; Li, X.; Li, Q.; Lyu, J.; Wang, W.; Zhuang, L.; Xu, Y. Magnetic drug-loaded osteoinductive Fe3O4/CaCO3 hybrid microspheres system: Efficient for sustained release of antibiotics. J. Phys. D Appl. Phys. 2020, 53, 245401. [Google Scholar] [CrossRef]
  85. Begines, B.; Ortiz, T.; Pérez-Aranda, M.; Martínez, G.; Merinero, M.; Argüelles-Arias, F.; Alcudia, A. Polymeric nanoparticles for drug delivery: Recent developments and future prospects. Nanomaterials 2020, 10, 1403. [Google Scholar] [CrossRef]
  86. Schwaminger, S.P.; Blank-Shim, S.A.; Scheifele, I.; Fraga-García, P.; Berensmeier, S. Peptide binding to metal oxide nanoparticles. Faraday Discuss. 2017, 204, 233–250. [Google Scholar] [CrossRef]
  87. Rauwolf, S.; Steegmüller, T.; Schwaminger, S.P.; Berensmeier, S. Purification of a peptide tagged protein via an affinity chromatographic process with underivatized silica. Eng. Life Sci. 2021, 21, 549–557. [Google Scholar] [CrossRef]
  88. Jiao, W.; Zhang, T.; Peng, M.; Yi, J.; He, Y.; Fan, H. Design of Magnetic Nanoplatforms for Cancer Theranostics. Biosensors 2022, 12, 38. [Google Scholar] [CrossRef]
  89. Gopalan Sibi, M.; Verma, D.; Kim, J. Magnetic core–shell nanocatalysts: Promising versatile catalysts for organic and photocatalytic reactions. Catal. Rev. Sci. Eng. 2020, 62, 163–311. [Google Scholar] [CrossRef]
  90. Zou, H.; Luo, Z.; Yang, X.; Xie, Q.; Zhou, Y. Toward emerging applications using core–shell nanostructured materials: A review. J. Mater. Sci. 2022, 57, 10912–10942. [Google Scholar] [CrossRef]
  91. Dinc, M.; Esen, C.; Mizaikoff, B. Recent advances on core–shell magnetic molecularly imprinted polymers for biomacromolecules. Trends Anal. Chem. 2019, 114, 202–217. [Google Scholar] [CrossRef]
  92. Tong, H.; Liu, X.J.; Zheng, M.S.; Dang, Z.M.; Zha, J.W. Dual functionalized Janus structural PVDF nanocomposite with surface-modified dielectric and magnetic nanoparticles. Appl. Phys. Lett. 2020, 117, 112903. [Google Scholar] [CrossRef]
  93. Zhang, X.; Fu, Q.; Duan, H.; Song, J.; Yang, H. Janus Nanoparticles: From Fabrication to (Bio)Applications. ACS Nano 2021, 15, 6147–6191. [Google Scholar] [CrossRef] [PubMed]
  94. Le, T.C.; Zhai, J.; Chiu, W.H.; Tran, P.A.; Tran, N. Janus particles: Recent advances in the biomedical applications. Int. J. Nanomed. 2019, 14, 6749–6777. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Pardo, A.; Gómez-Florit, M.; Barbosa, S.; Taboada, P.; Domingues, R.M.A.; Gomes, M.E. Magnetic Nanocomposite Hydrogels for Tissue Engineering: Design Concepts and Remote Actuation Strategies to Control Cell Fate. ACS Nano 2021, 15, 175–209. [Google Scholar] [CrossRef] [PubMed]
  96. Li, J.; Zhang, J.; Guo, Z.; Jiang, H.; Zhang, H.; Wang, X. Self-Assembly Fabrication of Honeycomb-like Magnetic-Fluorescent Fe3O4-QDs Nanocomposites for Bimodal Imaging. Langmuir 2020, 36, 14471–14477. [Google Scholar] [CrossRef] [PubMed]
  97. Gao, J.; Liang, G.; Cheung, J.S.; Pan, Y.; Kuang, Y.; Zhao, F.; Zhang, B.; Zhang, X.; Wu, E.X.; Xu, B. Multifunctional yolk-shell nanoparticles: A potential MRI contrast and anticancer agent. J. Am. Chem. Soc. 2008, 130, 11828–11833. [Google Scholar] [CrossRef]
  98. Mirbagheri, R.; Elhamifar, D.; Shaker, M. Yolk–shell structured magnetic mesoporous silica: A novel and highly efficient adsorbent for removal of methylene blue. Sci. Rep. 2021, 11, 1–15. [Google Scholar] [CrossRef]
  99. Díez, A.G.; Rincón-Iglesias, M.; Lanceros-Méndez, S.; Reguera, J.; Lizundia, E. Multicomponent magnetic nanoparticle engineering: The role of structure-property relationship in advanced applications. Mater. Today Chem. 2022, 26, 101220. [Google Scholar] [CrossRef]
  100. Zeng, Y.; Xu, G.; Kong, H.; Ye, G.; Guo, G.; Lu, C.; Nezamzadeh-Ejhieh, A.; Khan, M.S.; Liu, J.; Peng, Y. Recent advances of the core–shell MOFs in tumour therapy. Int. J. Pharm. 2022, 627, 122228. [Google Scholar] [CrossRef]
  101. Sanchez, L.M.; Alvarez, V.A. Advances in magnetic noble metal/iron-based oxide hybrid nanoparticles as biomedical devices. Bioeng. 2019, 6, 75. [Google Scholar] [CrossRef] [Green Version]
  102. Mourdikoudis, S.; Kostopoulou, A.; LaGrow, A.P. Magnetic Nanoparticle Composites: Synergistic Effects and Applications. Adv. Sci. 2021, 2004951, 1–57. [Google Scholar] [CrossRef]
  103. Ghosh, S.; Jiang, W.; McClements, J.D.; Xing, B. Colloidal stability of magnetic iron oxide nanoparticles: Influence of natural organic matter and synthetic polyelectrolytes. Langmuir 2011, 27, 8036–8043. [Google Scholar] [CrossRef]
  104. Darwish, M.S.A.; Al-Harbi, L.M.; Bakry, A. Synthesis of magnetite nanoparticles coated with polyvinyl alcohol for hyperthermia application. J. Therm. Anal. Calorim. 2022, 147, 11921–11930. [Google Scholar] [CrossRef]
  105. Zaloga, J.; Pöttler, M.; Leitinger, G.; Friedrich, R.P.; Almer, G.; Lyer, S.; Baum, E.; Tietze, R.; Heimke-Brinck, R.; Mangge, H.; et al. Pharmaceutical formulation of HSA hybrid coated iron oxide nanoparticles for magnetic drug targeting. Eur. J. Pharm. Biopharm. 2016, 101, 152–162. [Google Scholar] [CrossRef]
  106. Zaloga, J.; Stapf, M.; Nowak, J.; Pöttler, M.; Friedrich, R.P.; Tietze, R.; Lyer, S.; Lee, G.; Odenbach, S.; Hilger, I.; et al. Tangential flow ultrafiltration allows purification and concentration of lauric acid-/albumin-coated particles for improved magnetic treatment. Int. J. Mol. Sci. 2015, 16, 19291–19307. [Google Scholar] [CrossRef] [Green Version]
  107. Zaloga, J.; Janko, C.; Nowak, J.; Matuszak, J.; Knaup, S.; Eberbeck, D.; Tietze, R.; Unterweger, H.; Friedrich, R.P.; Duerr, S.; et al. Development of a lauric acid/albumin hybrid iron oxide nanoparticle system with improved biocompatibility. Int. J. Nanomed. 2014, 9, 4847–4866. [Google Scholar] [CrossRef] [Green Version]
  108. Corem-Salkmon, E.; Ram, Z.; Daniels, D.; Perlstein, B.; Last, D.; Salomon, S.; Tamar, G.; Shneor, R.; Guez, D.; Margel, S.; et al. Convection-enhanced delivery of methotrexate-loaded maghemite nanoparticles. Int. J. Nanomed. 2011, 6, 1595–1602. [Google Scholar] [CrossRef] [Green Version]
  109. Zhou, L.; Ye, L.; Lu, Y. Flexible and Effective Preparation of Magnetic Nanoclusters via One-Step Flow Synthesis. Nanomaterials 2022, 12, 350. [Google Scholar] [CrossRef]
  110. Junejo, Y.; Baykal, A.; Sözeri, H. Simple hydrothermal synthesis of Fe3O4-PEG nanocomposite. Cent. Eur. J. Chem. 2013, 11, 1527–1532. [Google Scholar] [CrossRef]
  111. Yallapu, M.M.; Foy, S.P.; Jain, T.K.; Labhasetwar, V. PEG-functionalized magnetic nanoparticles for drug delivery and magnetic resonance imaging applications. Pharm. Res. 2010, 27, 2283–2295. [Google Scholar] [CrossRef] [Green Version]
  112. Premaratne, G.; Coats, L.; Krishnan, S. NanoArmoring of Enzymes by Polymer-Functionalized Iron Oxide Nanoparticles, 1st ed.; Elsevier Inc.: Amsterdam, The Netherlands, 2017; Volume 590. [Google Scholar]
  113. Perera, A.S.; Zhang, S.; Homer-Vanniasinkam, S.; Coppens, M.O.; Edirisinghe, M. Polymer-Magnetic Composite Fibers for Remote-Controlled Drug Release. ACS Appl. Mater. Interfaces 2018, 10, 15524–15531. [Google Scholar] [CrossRef] [PubMed]
  114. Szymczyk, A.; Drozd, M.; Kamińska, A.; Matczuk, M.; Trzaskowski, M.; Mazurkiewicz-Pawlicka, M.; Ziółkowski, R.; Malinowska, E. Comparative Evaluation of Different Surface Coatings of Fe3O4-Based Magnetic Nano Sorbent for Applications in the Nucleic Acids Extraction. Int. J. Mol. Sci. 2022, 23, 8860. [Google Scholar] [CrossRef] [PubMed]
  115. Arias, L.S.; Pessan, J.P.; Vieira, A.P.M.; De Lima, T.M.T.; Delbem, A.C.B.; Monteiro, D.R. Iron oxide nanoparticles for biomedical applications: A perspective on synthesis, drugs, antimicrobial activity, and toxicity. Antibiotics 2018, 7, 46. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Ayub, A.; Wettig, S. An Overview of Nanotechnologies for Drug Delivery to the Brain. Pharmaceutics 2022, 14, 224. [Google Scholar] [CrossRef] [PubMed]
  117. Kadhim, W.K.A.; Nayef, U.M.; Jabir, M.S. Polyethylene glycol-functionalized magnetic (Fe3O4) nanoparticles: A good method for a successful antibacterial therapeutic agent via damage DNA molecule. Surf. Rev. Lett. 2019, 26, 1950079. [Google Scholar] [CrossRef]
  118. Jabir, M.S.; Nayef, U.M.; Kadhim, W.K.A. Polyethylene Glycol-Functionalized Magnetic (Fe3O4) Nanoparticles: A Novel DNA-Mediated Antibacterial Agent. Nano Biomed. Eng. 2019, 11, 18–27. [Google Scholar] [CrossRef]
  119. Gómez-Vallejo, V.; Puigivila, M.; Plaza-García, S.; Szczupak, B.; Piñol, R.; Murillo, J.L.; Sorribas, V.; Lou, G.; Veintemillas, S.; Ramos-Cabrer, P.; et al. PEG-copolymer-coated iron oxide nanoparticles that avoid the reticuloendothelial system and act as kidney MRI contrast agents. Nanoscale 2018, 10, 14153–14164. [Google Scholar] [CrossRef] [Green Version]
  120. Shen, L.; Li, B.; Qiao, Y. Fe3O4 nanoparticles in targeted drug/gene delivery systems. Materials 2018, 11, 324. [Google Scholar] [CrossRef] [Green Version]
  121. Ching, Y.C.; Gunathilake, T.M.S.U.; Chuah, C.H.; Ching, K.Y.; Singh, R.; Liou, N.S. Curcumin/Tween 20-incorporated cellulose nanoparticles with enhanced curcumin solubility for nano-drug delivery: Characterization and in vitro evaluation. Cellulose 2019, 26, 5467–5481. [Google Scholar] [CrossRef]
  122. Chen, Y.; Liu, Y.; Shi, Y.; Ping, J.; Wu, J.; Chen, H. Magnetic particles for integrated nucleic acid purification, amplification and detection without pipetting. TrAC Trends Anal. Chem. 2020, 127, 115912. [Google Scholar] [CrossRef]
  123. Godovikova, T.S.; Lisitskiy, V.A.; Antonova, N.M.; Popova, T.V.; Zakharova, O.D.; Chubarov, A.S.; Koptyug, I.V.; Sagdeev, R.Z.; Kaptein, R.; Akulov, A.E.; et al. Ligand-directed acid-sensitive amidophosphate 5-trifluoromethyl-2′-deoxyuridine conjugate as a potential theranostic agent. Bioconjug. Chem. 2013, 24, 780–795. [Google Scholar] [CrossRef]
  124. Khodadust, R.; Unal, O.; Acar, H.Y. Theranostic potential of self-luminescent branched polyethyleneimine-coated superparamagnetic iron oxide nanoparticles. Beilstein J. Nanotechnol. 2022, 13, 82–95. [Google Scholar] [CrossRef]
  125. Paris, J.L.; Vallet-Regí, M. Mesoporous silica nanoparticles for co-delivery of drugs and nucleic acids in oncology: A review. Pharmaceutics 2020, 12, 526. [Google Scholar] [CrossRef]
  126. Wang, R.; Degirmenci, V.; Xin, H.; Li, Y.; Wang, L.; Chen, J.; Hu, X.; Zhang, D. PEI-coated Fe3O4 nanoparticles enable efficient delivery of therapeutic siRNA targeting REST into glioblastoma cells. Int. J. Mol. Sci. 2018, 19, 2230. [Google Scholar] [CrossRef] [Green Version]
  127. Schneider, M.G.M.; Martín, M.J.; Otarola, J.; Vakarelska, E.; Simeonov, V.; Lassalle, V.; Nedyalkova, M. Biomedical Applications of Iron Oxide Nanoparticles: Current Insights Progress and Perspectives. Pharmaceutics 2022, 14, 204. [Google Scholar] [CrossRef]
  128. Raghava Reddy, K.; Reddy, P.A.; Reddy, C.V.; Shetti, N.P.; Babu, B.; Ravindranadh, K.; Shankar, M.V.; Reddy, M.C.; Soni, S.; Naveen, S. Functionalized Magnetic Nanoparticles/Biopolymer Hybrids: Synthesis Methods, Properties and Biomedical Applications, 1st ed.; Elsevier Ltd.: Amsterdam, The Netherlands, 2019; Volume 46, ISBN 9780128149928. [Google Scholar]
  129. Dmitrienko, E.V.; Pyshnaya, I.A.; Pyshnyi, D.V. Oligonucleotide Derivatives in the Hybridization Analysis of Nucleic Acids. I. Covalent Immobilization of Oligonucleotide Probes on Nylon. Russ. J. Bioorganic Chem. 2010, 36, 645–656. [Google Scholar] [CrossRef]
  130. Kolovskaya, O.S.; Zamay, T.N.; Zamay, G.S.; Babkin, V.A.; Medvedeva, E.N.; Neverova, N.A.; Kirichenko, A.K.; Zamay, S.S.; Lapin, I.N.; Morozov, E.V.; et al. Aptamer-conjugated superparamagnetic ferroarabinogalactan nanoparticles for targeted magnetodynamic therapy of cancer. Cancers 2020, 12, 216. [Google Scholar] [CrossRef] [Green Version]
  131. Thi, T.T.H.; Tran, D.H.N.; Bach, L.G.; Quang, H.V.; Nguyen, D.C.; Park, K.D.; Nguyen, D.H. Functional magnetic core-shell system-based iron oxide nanoparticle coated with biocompatible copolymer for anticancer drug delivery. Pharmaceutics 2019, 11, 120. [Google Scholar] [CrossRef] [Green Version]
  132. Piñeiro, Y.; Gómez, M.G.; Alves, L.d.C.; Prieto, A.A.; Acevedo, P.G.; Gudiña, R.S.; Puig, J.; Teijeiro, C.; Vilar, S.Y.; Rivas, J. Hybrid nanostructured magnetite nanoparticles: From bio-detection and theragnostics to regenerative medicine. Magnetochemistry 2020, 6, 4. [Google Scholar] [CrossRef] [Green Version]
  133. Li, Z.; Wang, Y.; Ni, Y.; Kokot, S. Fluorescence analysis of 6-mercaptopurine with the use of a nano-composite consisting of BSA-capped Au nano-clusters and core-shell Fe3O4-SiO2 nanoparticles. Biosens. Bioelectron. 2015, 70, 246–253. [Google Scholar] [CrossRef]
  134. Levy, I.; Sher, I.; Corem-Salkmon, E.; Ziv-Polat, O.; Meir, A.; Treves, A.J.; Nagler, A.; Kalter-Leibovici, O.; Margel, S.; Rotenstreich, Y. Bioactive magnetic near Infra-Red fluorescent core-shell iron oxide/human serum albumin nanoparticles for controlled release of growth factors for augmentation of human mesenchymal stem cell growth and differentiation. J. Nanobiotechnol. 2015, 13, 34. [Google Scholar] [CrossRef] [PubMed]
  135. Robinson, I.; Tung, L.D.; Maenosono, S.; Wälti, C.; Thanh, N.T.K. Synthesis of core-shell gold coated magnetic nanoparticles and their interaction with thiolated DNA. Nanoscale 2010, 2, 2624–2630. [Google Scholar] [CrossRef] [PubMed]
  136. Abarca-Cabrera, L.; Fraga-García, P.; Berensmeier, S. Bio-nano interactions: Binding proteins, polysaccharides, lipids and nucleic acids onto magnetic nanoparticles. Biomater. Res. 2021, 25, 1–18. [Google Scholar] [CrossRef]
  137. Uthaman, S.; Lee, S.J.; Cherukula, K.; Cho, C.S.; Park, I.K. Polysaccharide-coated magnetic nanoparticles for imaging and gene therapy. Biomed Res. Int. 2015, 2015, 14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Gan, W.; Gu, Y.; Han, J.; Li, C.X.; Sun, J.; Liu, P. Chitosan-Modified Filter Paper for Nucleic Acid Extraction and “in Situ PCR” on a Thermoplastic Microchip. Anal. Chem. 2017, 89, 3568–3575. [Google Scholar] [CrossRef]
  139. Samanta, B.; Yan, H.; Fischer, N.O.; Shi, J.; Jerry, D.J.; Rotello, V.M. Protein-passivated Fe3O4 nanoparticles: Low toxicity and rapid heating for thermal therapy. J. Mater. Chem. 2008, 18, 1204–1208. [Google Scholar] [CrossRef] [Green Version]
  140. Bychkova, A.V.; Sorokina, O.N.; Pronkin, P.G.; Tatikolov, A.S.; Kovarski, A.L.; Rosenfeld, M.A. Protein-Coated Magnetic Nanoparticles: Creation and Investigation. Proc. Int. Conf. Nanomater. Appl. Prop. 2013, 2, 1–5. [Google Scholar]
  141. Sakulkhu, U.; Mahmoudi, M.; Maurizi, L.; Salaklang, J.; Hofmann, H. Protein corona composition of superparamagnetic iron oxide nanoparticles with various physico-chemical properties and coatings. Sci. Rep. 2014, 4, 1–9. [Google Scholar] [CrossRef] [Green Version]
  142. Fouad, D.; Bachra, Y.; Ayoub, G.; Ouaket, A.; Bennamara, A.; Knouzi, N.; Berrada, M. A Novel Drug Delivery System Based on Nanoparticles of Magnetite Fe3O4 Embedded in an Auto Cross-Linked Chitosan. In Chitin and Chitosan—Physicochemical Properties and Industrial Applications; IntechOpen: London, UK, 2020; p. 290. ISBN 978-1-78984-425-2. [Google Scholar]
  143. Bychkova, A.V.; Yakunina, M.N.; Lopukhova, M.V.; Degtyarev, Y.N.; Motyakin, M.V.; Pokrovsky, V.S.; Kovarski, A.L.; Gorobets, M.G.; Retivov, V.M.; Khachatryan, D.S. Albumin-Functionalized Iron Oxide Nanoparticles for Theranostics: Engineering and Long-Term In Situ Imaging. Pharmaceutics 2022, 14, 2771. [Google Scholar] [CrossRef]
  144. Chuang, V.T.G.; Maruyama, T.; Otagiri, M. Human Serum Albumin in Blood Detoxification Treatment. In Albumin in Medicine; Springer Singapore: Singapore, 2016; pp. 209–225. [Google Scholar]
  145. Kragh-hansen, U. Human Serum Albumin: A Multifunctional Protein. In Albumine in Medicine; Springer Singapore: Singapore, 2016; pp. 1–24. ISBN 978-981-10-2115-2. [Google Scholar]
  146. Fanali, G.; di Masi, A.; Trezza, V.; Marino, M.; Fasano, M.; Ascenzi, P. Human serum albumin: From bench to bedside. Mol. Asp. Med. 2012, 33, 209–290. [Google Scholar] [CrossRef]
  147. Li, H.; Wang, Y.; Tang, Q.; Yin, D.; Tang, C.; He, E.; Zou, L.; Peng, Q. The Protein Corona and its Effects on Nanoparticle-Based Drug Delivery Systems. Acta Biomater. 2021, 129, 57–72. [Google Scholar] [CrossRef]
  148. Baki, A.; Remmo, A.; Löwa, N.; Wiekhorst, F.; Bleul, R. Albumin-coated single-core iron oxide nanoparticles for enhanced molecular magnetic imaging (Mri/mpi). Int. J. Mol. Sci. 2021, 22, 6235. [Google Scholar] [CrossRef]
  149. Rahdar, S.; Rahdar, A.; Ahmadi, S.; Trant, J.F. Adsorption of bovine serum albumin (BSA) by bare magnetite nanoparticles with surface oxidative impurities that prevent aggregation. Can. J. Chem. 2019, 97, 577–583. [Google Scholar] [CrossRef] [Green Version]
  150. Ziegler-Borowska, M. Magnetic nanoparticles coated with aminated starch for HSA immobilization- simple and fast polymer surface functionalization. Int. J. Biol. Macromol. 2019, 136, 106–114. [Google Scholar] [CrossRef]
  151. Moya, C.; Escudero, R.; Malaspina, D.C.; De La Mata, M.; Hernández-Saz, J.; Faraudo, J.; Roig, A. Insights into Preformed Human Serum Albumin Corona on Iron Oxide Nanoparticles: Structure, Effect of Particle Size, Impact on MRI Efficiency, and Metabolization. ACS Appl. Bio Mater. 2019, 2, 3084–3094. [Google Scholar] [CrossRef]
  152. Mariam, J.; Sivakami, S.; Dongre, P.M. Albumin corona on nanoparticles–a strategic approach in drug delivery. Drug Deliv. 2016, 23, 2668–2676. [Google Scholar] [CrossRef] [Green Version]
  153. Hassanin, I.; Elzoghby, A. Albumin-based nanoparticles: A promising strategy to overcome cancer drug resistance. Cancer Drug Resist. 2020, 3, 930–946. [Google Scholar] [CrossRef]
  154. Srivastava, A.; Prajapati, A. Albumin and functionalized albumin nanoparticles: Production strategies, characterization, and target indications. Asian Biomed. 2020, 14, 217–242. [Google Scholar] [CrossRef]
  155. Bolaños, K.; Kogan, M.J.; Araya, E. Capping gold nanoparticles with albumin to improve their biomedical properties. Int. J. Nanomed. 2019, 14, 6387–6406. [Google Scholar] [CrossRef] [Green Version]
  156. Popova, T.V.; Khan, H.; Chubarov, A.S.; Lisitskiy, V.A.; Antonova, N.M.; Akulov, A.E.; Shevelev, O.B.; Zavjalov, E.L.; Silnikov, V.N.; Ahmad, S.; et al. Biotin-decorated anti-cancer nucleotide theranostic conjugate of human serum albumin: Where the seed meets the soil? Bioorganic Med. Chem. Lett. 2018, 28, 260–264. [Google Scholar] [CrossRef]
  157. Erdal, E.; Demirbilek, M.; Yeh, Y.; Akbal, Ö.; Ruff, L.; Bozkurt, D.; Cabuk, A.; Senel, Y.; Gumuskaya, B.; Algın, O.; et al. A Comparative Study of Receptor-Targeted Magnetosome and HSA-Coated Iron Oxide Nanoparticles as MRI Contrast-Enhancing Agent in Animal Cancer Model. Appl. Biochem. Biotechnol. 2018, 185, 91–113. [Google Scholar] [CrossRef] [PubMed]
  158. Abakumov, M.A.; Nukolova, N.V.; Sokolsky-Papkov, M.; Shein, S.A.; Sandalova, T.O.; Vishwasrao, H.M.; Grinenko, N.F.; Gubsky, I.L.; Abakumov, A.M.; Kabanov, A.V.; et al. VEGF-targeted magnetic nanoparticles for MRI visualization of brain tumor. Nanomed. Nanotechnol. Biol. Med. 2015, 11, 825–833. [Google Scholar] [CrossRef] [PubMed]
  159. Chubarov, A.S.; Shakirov, M.M.; Koptyug, I.V.; Sagdeev, R.Z.; Knorre, D.G.; Godovikova, T.S. Synthesis and characterization of fluorinated homocysteine derivatives as potential molecular probes for 19F magnetic resonance spectroscopy and imaging. Bioorg. Med. Chem. Lett. 2011, 21, 4050–4053. [Google Scholar] [CrossRef] [PubMed]
  160. Chubarov, A.S.; Zakharova, O.D.; Koval, O.A.; Romaschenko, A.V.; Akulov, A.E.; Zavjalov, E.L.; Razumov, I.A.; Koptyug, I.V.; Knorre, D.G.; Godovikova, T.S. Design of protein homocystamides with enhanced tumor uptake properties for 19F magnetic resonance imaging. Bioorg. Med. Chem. 2015, 23, 6943–6954. [Google Scholar] [CrossRef] [PubMed]
  161. Lisitskiy, V.A.; Khan, H.; Popova, T.V.; Chubarov, A.S.; Zakharova, O.D.; Akulov, A.E.; Shevelev, O.B.; Zavjalov, E.L.; Koptyug, I.V.; Moshkin, M.P.; et al. Multifunctional human serum albumin-therapeutic nucleotide conjugate with redox and pH-sensitive drug release mechanism for cancer theranostics. Bioorganic Med. Chem. Lett. 2017, 27, 3925–3930. [Google Scholar] [CrossRef]
  162. Dobrynin, S.; Kutseikin, S.; Morozov, D.; Krumkacheva, O.; Spitsyna, A.; Gatilov, Y.; Silnikov, V.; Angelovski, G.; Bowman, M.K.; Kirilyuk, I.; et al. Human Serum Albumin Labelled with Sterically-Hindered Nitroxides as Potential MRI Contrast Agents. Molecules 2020, 25, 1709. [Google Scholar] [CrossRef] [Green Version]
  163. Hou, X.; Zhang, H.; Li, H.; Zhang, D. Magnetic albumin immuno-nanospheres as an efficient gene delivery system for a potential use in lung cancer: Preparation, in vitro targeting and biological effect analysis. J. Drug Target. 2016, 24, 247–256. [Google Scholar] [CrossRef]
  164. Prajapati, R.; Somoza, Á. Albumin nanostructures for nucleic acid delivery in cancer: Current trend, emerging issues, and possible solutions. Cancers 2021, 13, 3454. [Google Scholar] [CrossRef]
  165. Ball, V. Polydopamine nanomaterials: Recent advances in synthesis methods and applications. Front. Bioeng. Biotechnol. 2018, 6, 109. [Google Scholar] [CrossRef] [Green Version]
  166. Li, M.; Wang, Y.; Han, X.; Liu, Y.; Ma, M.; Zhang, L. Multifunctional Polydopamine-Based Nanoparticles for Dual-Mode Imaging Guided Targeted Therapy of Lupus Nephritis. Pharmaceutics 2022, 14, 1988. [Google Scholar] [CrossRef]
  167. García Acevedo, P.; González Gómez, M.A.; Arnosa Prieto, Á.; De Castro Alves, L.; Seco Gudiña, R.; Piñeiro, Y.; Rivas, J. Fluorescent Single-Core and Multi-Core Nanoprobes as Cell Trackers and Magnetic Nanoheaters. Magnetochemistry 2022, 8, 83. [Google Scholar] [CrossRef]
  168. Niezni, D.; Harris, Y.; Sason, H.; Avrashami, M.; Shamay, Y. Polydopamine Copolymers for Stable Drug Nanoprecipitation. Int. J. Mol. Sci. 2022, 23, 12420. [Google Scholar] [CrossRef]
  169. Zandieh, M.; Liu, J. Spherical Nucleic Acid Mediated Functionalization of Polydopamine-Coated Nanoparticles for Selective DNA Extraction and Detection. Bioconjug. Chem. 2021, 32, 801–809. [Google Scholar] [CrossRef]
  170. Singh, I.; Dhawan, G.; Gupta, S.; Kumar, P. Recent Advances in a Polydopamine-Mediated Antimicrobial Adhesion System. Front. Microbiol. 2021, 11, 607099. [Google Scholar] [CrossRef]
  171. Zhang, M.; Li, L.; Li, B.; Tian, N.; Yang, M.; Zhang, H.; You, C.; Zhang, J. Adsorption of DNA by using polydopamine modified magnetic nanoparticles based on solid-phase extraction. Anal. Biochem. 2019, 579, 9–17. [Google Scholar] [CrossRef]
  172. Zhao, Z.; Cui, H.; Song, W.; Ru, X.; Zhou, W.; Yu, X. A simple magnetic nanoparticles-based viral RNA extraction method for efficient detection of SARS-CoV-2. bioRxiv 2020. biorXiv:518055.2020.02.22.961268. [Google Scholar]
  173. Spoială, A.; Ilie, C.-I.; Crăciun, L.N.; Ficai, D.; Ficai, A.; Andronescu, E. Magnetite-Silica Core/Shell Nanostructures: From Surface Functionalization towards Biomedical Applications—A Review. Appl. Sci. 2021, 11, 11075. [Google Scholar] [CrossRef]
  174. Yue, Q.; Sun, J.; Kang, Y.; Deng, Y. Advances in the Interfacial Assembly of Mesoporous Silica on Magnetite Particles. Angew. Chem. 2020, 132, 15936–15949. [Google Scholar] [CrossRef]
  175. Bagheri, E.; Ansari, L.; Abnous, K.; Taghdisi, S.M.; Charbgoo, F.; Ramezani, M.; Alibolandi, M. Silica based hybrid materials for drug delivery and bioimaging. J. Control. Release 2018, 277, 57–76. [Google Scholar] [CrossRef]
  176. Frickenstein, A.N.; Hagood, J.M.; Britten, C.N.; Abbott, B.S.; McNally, M.W.; Vopat, C.A.; Patterson, E.G.; Maccuaig, W.M.; Jain, A.; Walters, K.B.; et al. Mesoporous silica nanoparticles: Properties and strategies for enhancing clinical effect. Pharmaceutics 2021, 13, 570. [Google Scholar] [CrossRef]
  177. Zhu, N.; Ji, H.; Yu, P.; Niu, J.; Farooq, M.U.; Akram, M.W.; Udego, I.O.; Li, H.; Niu, X. Surface modification of magnetic iron oxide nanoparticles. Nanomaterials 2018, 8, 810. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  178. Husain, H.; Hariyanto, B.; Sulthonul, M.; Klysubun, W.; Darminto; Pratapa, S. Structure and magnetic properties of silica-coated magnetite- nanoparticle composites. Mater. Res. Express 2019, 6, 086117. [Google Scholar] [CrossRef]
  179. Corma, A.; Botella, P.; Rivero-Buceta, E. Silica-Based Stimuli-Responsive Systems for Antitumor Drug Delivery and Controlled Release. Pharmaceutics 2022, 14, 110. [Google Scholar] [CrossRef] [PubMed]
  180. Pontón, I.; del Rio, A.M.; Gómez, M.G.; Sánchez-García, D. Preparation and applications of organo-silica hybrid mesoporous silica nanoparticles for the co-delivery of drugs and nucleic acids. Nanomaterials 2020, 10, 2466. [Google Scholar] [CrossRef] [PubMed]
  181. Schwaminger, S.P.; Fraga-García, P.; Eigenfeld, M.; Becker, T.M.; Berensmeier, S. Magnetic separation in bioprocessing beyond the analytical scale: From biotechnology to the food industry. Front. Bioeng. Biotechnol. 2019, 7, 233. [Google Scholar] [CrossRef] [PubMed]
  182. Ahmad, T.; Bae, H.; Rhee, I.; Chang, Y.; Jin, S.U.; Hong, S. Gold-coated iron oxide nanoparticles as a T2 agent in magnetic resonance imaging. J. Nanosci. Nanotechnol. 2012, 12, 5132–5137. [Google Scholar] [CrossRef] [Green Version]
  183. Efremova, M.V.; Naumenko, V.A.; Spasova, M.; Garanina, A.S.; Abakumov, M.A.; Blokhina, A.D.; Melnikov, P.A.; Prelovskaya, A.O.; Heidelmann, M.; Li, Z.A.; et al. Magnetite-Gold nanohybrids as ideal all-in-one platforms for theranostics. Sci. Rep. 2018, 8, 1–19. [Google Scholar] [CrossRef] [Green Version]
  184. Du, B.W.; Chu, C.Y.; Lin, C.C.; Ko, F.H. The multifunctionally graded system for a controlled size effect on iron oxide–gold based core-shell nanoparticles. Nanomaterials 2021, 11, 1695. [Google Scholar] [CrossRef]
  185. Żuk, M.; Podgórski, R.; Ruszczyńska, A.; Ciach, T.; Majkowska-Pilip, A.; Bilewicz, A.; Krysiński, P. Multifunctional Nanoparticles Based on Iron Oxide and Gold-198 Designed for Magnetic Hyperthermia and Radionuclide Therapy as a Potential Tool for Combined HER2-Positive Cancer Treatment. Pharmaceutics 2022, 14, 1680. [Google Scholar] [CrossRef]
  186. Abdulkadhim, W.K.; Kut, A. A New DNA-Mediated Antibacterial Agent Magnetic (Fe3O4) Nanoparticles with Gold and Silver Functionalization. Wasit J. Pure Sci. 2022, 1, 248–259. [Google Scholar] [CrossRef]
  187. Pang, Y.; Wang, C.; Wang, J.; Sun, Z.; Xiao, R.; Wang, S. Fe3O4@Ag magnetic nanoparticles for microRNA capture and duplex-specific nuclease signal amplification based SERS detection in cancer cells. Biosens. Bioelectron. 2016, 79, 574–580. [Google Scholar] [CrossRef]
  188. Nguyen-Tri, P.; Nguyen, V.T.; Nguyen, T.A. Biological activity and nanostructuration of Fe3O4-Ag/high density polyethylene nanocomposites. J. Compos. Sci. 2019, 3, 34. [Google Scholar] [CrossRef] [Green Version]
  189. Tintoré, M.; Mazzini, S.; Polito, L.; Marelli, M.; Latorre, A.; Somoza, Á.; Aviñó, A.; Fàbrega, C.; Eritja, R. Gold-coated superparamagnetic nanoparticles for single methyl discrimination in DNA aptamers. Int. J. Mol. Sci. 2015, 16, 27625–27639. [Google Scholar] [CrossRef]
  190. León Félix, L.; Sanz, B.; Sebastián, V.; Torres, T.E.; Sousa, M.H.; Coaquira, J.A.H.; Ibarra, M.R.; Goya, G.F. Gold-decorated magnetic nanoparticles design for hyperthermia applications and as a potential platform for their surface-functionalization. Sci. Rep. 2019, 9, 1–11. [Google Scholar] [CrossRef] [Green Version]
  191. Oza, G.; Krishnajyothi, K.; Merupo, V.I.; Bracamontes, K.A.C.; Olmos, P.C.; Garrido, E.; Velumani, S.; Sridharan, M.; Sharma, A.; Arriaga, L.G.; et al. Gold-Iron oxide yolk-shell nanoparticles (YSNPs) as magnetic probe for fluorescence-based detection of 3 base mismatch DNA. Colloids Surf. B Biointerfaces 2019, 176, 431–438. [Google Scholar] [CrossRef]
  192. Lee, M.H.; Leu, C.C.; Lin, C.C.; Tseng, Y.F.; Lin, H.Y.; Yang, C.N. Gold-decorated magnetic nanoparticles modified with hairpin-shaped DNA for fluorometric discrimination of single-base mismatch DNA. Microchim. Acta 2019, 186, 1–8. [Google Scholar] [CrossRef]
  193. Xuhong, Y.; Sinong, Z.; Jianping, L.; Yu, C.; Juanli, Z.; Chao, Z.; Desheng, L.; Kai, H.; Yali, C.; Wenli, H. A PCR-lateral flow assay system based on gold magnetic nanoparticles for CYP2C19 genotyping and its clinical applications. Artif. Cells Nanomed. Biotechnol. 2019, 47, 636–643. [Google Scholar] [CrossRef] [Green Version]
  194. Karami, F.; Noori-Daloii, M.R.; Omidfar, K.; Tabrizi, M.; Hantooshzadeh, S.; Aleyasin, A.; Daneshpour, M.; Modarressi, M.H. Modified methylated DNA immunoprecipitation protocol for noninvasive prenatal diagnosis of Down syndrome. J. Obstet. Gynaecol. Res. 2018, 44, 608–613. [Google Scholar] [CrossRef]
  195. Epanchintseva, A.V.; Gorbunova, E.A.; Ryabchikova, E.I.; Pyshnaya, I.A.; Pyshnyi, D.V. Effect of Fluorescent Labels on DNA Affinity for Gold Nanoparticles. Nanomaterials 2021, 11, 1178. [Google Scholar] [CrossRef]
  196. Abedin, M.R.; Umapathi, S.; Mahendrakar, H.; Laemthong, T.; Coleman, H.; Muchangi, D.; Santra, S.; Nath, M.; Barua, S. Polymer coated gold-ferric oxide superparamagnetic nanoparticles for theranostic applications. J. Nanobiotechnol. 2018, 16, 1–13. [Google Scholar] [CrossRef] [Green Version]
  197. Alromi, D.A.; Madani, S.Y.; Seifalian, A. Emerging application of magnetic nanoparticles for diagnosis and treatment of cancer. Polymers 2021, 13, 4146. [Google Scholar] [CrossRef] [PubMed]
  198. Han, C.; Zhang, A.; Kong, Y.; Yu, N.; Xie, T.; Dou, B.; Li, K.; Wang, Y.; Li, J.; Xu, K. Multifunctional iron oxide-carbon hybrid nanoparticles for targeted fluorescent/MR dual-modal imaging and detection of breast cancer cells. Anal. Chim. Acta 2019, 1067, 115–128. [Google Scholar] [CrossRef] [PubMed]
  199. Wu, H.; Huang, Q.; Tan, Y. Carbon nanomaterials for biomedical applications. In Carbon Nanomaterials; CRC: Boca Raton, FL, USA, 2019; pp. 255–293. [Google Scholar] [CrossRef]
  200. Yoon, J.; Shin, M.; Lee, T.; Choi, J.W. Highly sensitive biosensors based on biomolecules and functional nanomaterials depending on the types of nanomaterials: A perspective review. Materials 2020, 13, 299. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  201. Porras, J.C.; Bernuz, M.; Marfa, J.; Pallares-Rusiñol, A.; Martí, M.; Pividori, M.I. Comparative study of gold and carbon nanoparticles in nucleic acid lateral flow assay. Nanomaterials 2021, 11, 741. [Google Scholar] [CrossRef]
  202. Siddiqui, M.T.H.; Nizamuddin, S.; Baloch, H.A.; Mubarak, N.M.; Al-Ali, M.; Mazari, S.A.; Bhutto, A.W.; Abro, R.; Srinivasan, M.; Griffin, G. Fabrication of advance magnetic carbon nano-materials and their potential applications: A review. J. Environ. Chem. Eng. 2019, 7, 102812. [Google Scholar] [CrossRef]
  203. Vázquez-González, M.; Willner, I. Aptamer-functionalized hybrid nanostructures for sensing, drug delivery, catalysis and mechanical applications. Int. J. Mol. Sci. 2021, 22, 1803. [Google Scholar] [CrossRef]
  204. Trofimov, A.D.; Ivanova, A.A.; Zyuzin, M.V.; Timin, A.S. Porous Inorganic Carriers based on Silica, Calcium Carbonate and Calcium Phosphate for Controlled/Modulated Drug Delivery: Fresh Outlook and Future Perspectives. Pharmaceutics 2018, 10, 167. [Google Scholar] [CrossRef] [Green Version]
  205. Zhao, P.; Tian, Y.; You, J.; Hu, X. Recent Advances of Calcium Carbonate Nanoparticles for Biomedical Applications. Bioengineering 2022, 9, 691. [Google Scholar] [CrossRef]
  206. Popova, V.; Poletaeva, Y.; Pyshnaya, I.; Pyshnyi, D.; Dmitrienko, E. Designing pH-Dependent Systems Based on Nanoscale Calcium Carbonate for the Delivery of an Antitumor Drug. Nanomaterials 2021, 11, 2794. [Google Scholar] [CrossRef]
  207. Luo, W.; Hua, J.; Xie, X. Polyethylenimine-CO2 adduct-stabilized vaterite hydrocolloidal particles. Mater. Chem. Phys. 2023, 294, 127025. [Google Scholar] [CrossRef]
  208. Persano, F.; Nobile, C.; Piccirillo, C.; Gigli, G.; Leporatti, S. Monodisperse and Nanometric-Sized Calcium Carbonate Particles Synthesis Optimization. Nanomaterials 2022, 12, 1494. [Google Scholar] [CrossRef]
  209. Atchudan, R.; Perumal, S.; Joo, J.; Lee, Y.R. Synthesis and Characterization of Monodispersed Spherical Calcium Oxide and Calcium Carbonate Nanoparticles via Simple Pyrolysis. Nanomaterials 2022, 12, 2424. [Google Scholar] [CrossRef]
  210. He, X.W.; Liu, T.; Chen, Y.X.; Cheng, D.J.; Li, X.R.; Xiao, Y.; Feng, Y.L. Calcium carbonate nanoparticle delivering vascular endothelial growth factor-C siRNA effectively inhibits lymphangiogenesis and growth of gastric cancer in vivo. Cancer Gene Ther. 2008, 15, 193–202. [Google Scholar] [CrossRef]
  211. Maleki Dizaj, S.; Sharifi, S.; Ahmadian, E.; Eftekhari, A.; Adibkia, K.; Lotfipour, F. An update on calcium carbonate nanoparticles as cancer drug/gene delivery system. Expert Opin. Drug Deliv. 2019, 16, 331–345. [Google Scholar] [CrossRef]
  212. Zhang, W.; Li, Q.; Li, J.; Sun, X.; Shen, J.; Han, W.; Wang, L. The preparation of layered hierarchical and cube-shaped magnetic Fe3O4/CaCO3 for efficient enrichment of Pb(Ⅱ) from aqueous solutions. Environ. Nanotechnol. Monit. Manag. 2021, 16, 100600. [Google Scholar] [CrossRef]
  213. Serov, N.; Prilepskii, A.; Sokolov, A.; Vinogradov, V. Synthesis of Plasmin-Loaded Fe3O4@CaCO3 Nanoparticles: Towards Next-Generation Thrombolytic Drugs. ChemNanoMat 2019, 5, 1267–1271. [Google Scholar] [CrossRef]
  214. Li, F.H.; Tang, N.; Wang, Y.Q.; Zhang, L.; Du, W.; Xiang, J.; Cheng, P.G. Synthesis and Characterization of Magnetic Carriers Based on Immobilized Enzyme. IOP Conf. Ser. Mater. Sci. Eng. 2018, 359, 012044. [Google Scholar] [CrossRef] [Green Version]
  215. Lee, Y.H.; Seo, J.C.; Oh, Y.K.; Lee, K. Synthesis of microaglae-capturing magnetic microcapsule using CaCO3 microparticles and layer-by-layer coating. Korean J. Mater. Res. 2018, 28, 376–380. [Google Scholar] [CrossRef]
  216. Han, P.; Jiang, Z.; Wang, X.; Wang, X.; Zhang, S.; Shi, J.; Wu, H. Facile preparation of porous magnetic polydopamine microspheres through an inverse replication strategy for efficient enzyme immobilization. J. Mater. Chem. B 2015, 3, 7194–7202. [Google Scholar] [CrossRef]
  217. Wang, C.; Yan, J.; Cui, X.; Cong, D.; Wang, H. Preparation and characterization of magnetic hollow PMMA nanospheres via in situ emulsion polymerization. Colloids Surf. A Physicochem. Eng. Asp. 2010, 363, 71–77. [Google Scholar] [CrossRef]
  218. Ma, H.; Zhou, J.; Caruntu, D.; Yu, M.H.; Chen, J.F.; O’Connor, C.J.; Zhou, W.L. Fabrication of magnetic porous hollow silica drug carriers using CaCO3 Fe3O4 composite nanoparticles and cationic surfactant double templates. J. Appl. Phys. 2008, 103, 07A320. [Google Scholar] [CrossRef]
  219. Liu, B.; Jiang, M.; Zhu, D.; Zhang, J.; Wei, G. Metal-organic frameworks functionalized with nucleic acids and amino acids for structure- and function-specific applications: A tutorial review. Chem. Eng. J. 2022, 428, 131118. [Google Scholar] [CrossRef]
  220. Zhuang, J.; Young, A.P.; Tsung, C.K. Integration of Biomolecules with Metal–Organic Frameworks. Small 2017, 13, 1700880. [Google Scholar] [CrossRef] [PubMed]
  221. Zhao, X.; Liu, S.; Tang, Z.; Niu, H.; Cai, Y.; Meng, W.; Wu, F.; Giesy, J.P. Synthesis of magnetic metal-organic framework (MOF) for efficient removal of organic dyes from water. Sci. Rep. 2015, 5, 1–10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  222. Ma, M.; Lu, X.; Guo, Y.; Wang, L.; Liang, X. Combination of metal-organic frameworks (MOFs) and covalent organic frameworks (COFs): Recent advances in synthesis and analytical applications of MOF/COF composites. Trends Anal. Chem. 2022, 157, 116741. [Google Scholar] [CrossRef]
  223. Malhotra, N.; Lee, J.S.; Liman, R.A.D.; Ruallo, J.M.S.; Villaflore, O.B.; Ger, T.R.; Hsiao, C. Der Potential toxicity of iron oxide magnetic nanoparticles: A review. Molecules 2020, 25, 3159. [Google Scholar] [CrossRef]
  224. Chrishtop, V.V.; Mironov, V.A.; Prilepskii, A.Y.; Nikonorova, V.G.; Vinogradov, V.V. Organ-specific toxicity of magnetic iron oxide-based nanoparticles. Nanotoxicology 2021, 15, 167–204. [Google Scholar] [CrossRef]
  225. Mosmann, T. Rapid colorimetric assay for cellular growth and survival: Application to proliferation and cytotoxicity assays. J. Immunol. Methods 1983, 65, 55–63. [Google Scholar] [CrossRef]
  226. Präbst, K.; Engelhardt, H.; Ringgeler, S.; Hübner, H. Basic Colorimetric Proliferation Assays: MTT, WST, and Resazurin. In Cell Viability Assays. Methods in Molecular Biology; Springer: Cham, Switzerland, 2017; pp. 1–17. [Google Scholar]
  227. Attarilar, S.; Yang, J.; Ebrahimi, M.; Wang, Q.; Liu, J.; Tang, Y.; Yang, J. The Toxicity Phenomenon and the Related Occurrence in Metal and Metal Oxide Nanoparticles: A Brief Review From the Biomedical Perspective. Front. Bioeng. Biotechnol. 2020, 8, 822. [Google Scholar] [CrossRef]
  228. Liu, G.; Gao, J.; Ai, H.; Chen, X. Applications and potential toxicity of magnetic iron oxide nanoparticles. Small 2013, 9, 1533–1545. [Google Scholar] [CrossRef]
  229. Abakumov, M.A.; Semkina, A.S.; Skorikov, A.S.; Vishnevskiy, D.A.; Ivanova, A.V.; Mironova, E.; Davydova, G.A.; Majouga, A.G.; Chekhonin, V.P. Toxicity of iron oxide nanoparticles: Size and coating effects. J. Biochem. Mol. Toxicol. 2018, 32, e22225. [Google Scholar] [CrossRef]
  230. De Simone, U.; Roccio, M.; Gribaldo, L.; Spinillo, A.; Caloni, F.; Coccini, T. Human 3D cultures as models for evaluating magnetic nanoparticle CNS cytotoxicity after short- and repeated long-term exposure. Int. J. Mol. Sci. 2018, 19, 1993. [Google Scholar] [CrossRef]
  231. Kappes, M.; Friedrich, B.; Pfister, F.; Huber, C.; Friedrich, R.P.; Stein, R.; Braun, C.; Band, J.; Schreiber, E.; Alexiou, C.; et al. Superparamagnetic Iron Oxide Nanoparticles for Targeted Cell Seeding: Magnetic Patterning and Magnetic 3D Cell Culture. Adv. Funct. Mater. 2022, 32, 2203672. [Google Scholar] [CrossRef]
  232. Nguyen, K.; Nuß, B.; Mühlberger, M.; Unterweger, H.; Friedrich, R.P.; Alexiou, C.; Janko, C. Superparamagnetic iron oxide nanoparticles carrying chemotherapeutics improve drug efficacy in monolayer and spheroid cell culture by enabling active accumulation. Nanomaterials 2020, 10, 1577. [Google Scholar] [CrossRef]
  233. Anisimov, R.A.; Gorin, D.A.; Abalymov, A.A. 3D Cell Spheroids as A Tool for Evaluating the Effectiveness of Carbon Nanotubes as A Drug Delivery and Photothermal Therapy Agents. C 2022, 8, 56. [Google Scholar] [CrossRef]
  234. Henrique, R.B.L.; Lima, R.R.M.; Monteiro, C.A.P.; Oliveira, W.F.; Pereira, G.; Cabral Filho, P.E.; Fontes, A. Advances in the study of spheroids as versatile models to evaluate biological interactions of inorganic nanoparticles. Life Sci. 2022, 302, 120657. [Google Scholar] [CrossRef]
  235. Juarez-Moreno, K.; Chávez-García, D.; Hirata, G.; Vazquez-Duhalt, R. Monolayer (2D) or spheroids (3D) cell cultures for nanotoxicological studies? Comparison of cytotoxicity and cell internalization of nanoparticles. Toxicol. Vitr. 2022, 85, 105461. [Google Scholar] [CrossRef]
  236. Canaparo, R.; Foglietta, F.; Limongi, T.; Serpe, L. Biomedical applications of reactive oxygen species generation by metal nanoparticles. Materials 2021, 14, 53. [Google Scholar] [CrossRef]
  237. Nelson, N.; Port, J.; Pandey, M. Use of Superparamagnetic Iron Oxide Nanoparticles (SPIONs) via Multiple Imaging Modalities and Modifications to Reduce Cytotoxicity: An Educational Review. J. Nanotheranostics 2020, 1, 105–135. [Google Scholar] [CrossRef]
  238. Laurent, N.; Sapet, C.; Gourrierec, L.L.; Bertosio, E.; Zelphati, O. Nucleic acid delivery using magnetic nanoparticles: The MagnetofectionTM technology. Ther. Deliv. 2011, 2, 471–482. [Google Scholar] [CrossRef]
  239. Kami, D.; Takeda, S.; Itakura, Y.; Gojo, S.; Watanabe, M.; Toyoda, M. Application of magnetic nanoparticles to gene delivery. Int. J. Mol. Sci. 2011, 12, 3705–3722. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  240. Bakshi, S.; Zakharchenko, A.; Minko, S.; Kolpashchikov, D.; Katz, E. Towards Nanomaterials for Cancer Theranostics: A System of DNA-Modified Magnetic Nanoparticles for Detection and Suppression of RNA Marker in Cancer Cells. Magnetochemistry 2019, 5, 24. [Google Scholar] [CrossRef]
  241. Kenchegowda, M.; Rahamathulla, M.; Hani, U.; Begum, M.Y.; Guruswamy, S.; Osmani, R.A.M.; Gowrav, M.P.; Alshehri, S.; Ghoneim, M.M.; Alshlowi, A.; et al. Smart Nanocarriers as an Emerging Platform for Cancer Therapy: A Review. Molecules 2022, 27, 146. [Google Scholar] [CrossRef]
  242. Leach, J.C.; Wang, A.; Ye, K.; Jin, S. A RNA-DNA hybrid aptamer for nanoparticle-based prostate tumor targeted drug delivery. Int. J. Mol. Sci. 2016, 17, 380. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  243. Taghavi Pourianazar, N.; Gunduz, U. CpG oligodeoxynucleotide-loaded PAMAM dendrimer-coated magnetic nanoparticles promote apoptosis in breast cancer cells. Biomed. Pharmacother. 2016, 78, 81–91. [Google Scholar] [CrossRef]
  244. Bassetto, M.; Sen, M.; Poulhes, F.; Arango-Gonzalez, B.; Bonvin, E.; Sapet, C.; Ueffing, M.; Zelphati, O. New Method for Efficient siRNA Delivery in Retina Explants: Reverse Magnetofection. Bioconjug. Chem. 2021, 32, 1078–1093. [Google Scholar] [CrossRef]
  245. Gozuacik, D.; Akkoc, Y.; Kosar, A.; Dogan-Ekici, A.I.; Ekici, S. Anticancer Use of Nanoparticles as Nucleic Acid Carriers. J. Biomed. Nanotechnol. 2014, 10, 1751–1783. [Google Scholar] [CrossRef]
  246. Kostevšek, N. A review on the optimal design of magnetic nanoparticle-based t2 mri contrast agents. Magnetochemistry 2020, 6, 11. [Google Scholar] [CrossRef] [Green Version]
  247. Wallyn, J.; Anton, N.; Vandamme, T.F. Synthesis, principles, and properties of magnetite nanoparticles for in vivo imaging applications—A review. Pharmaceutics 2019, 11, 601. [Google Scholar] [CrossRef] [Green Version]
  248. Ellis, C.M.; Pellico, J.; Davis, J.J. Magnetic Nanoparticles Supporting Bio-responsive T1/T2 Magnetic Resonance Imaging. Materials 2019, 12, 4096. [Google Scholar] [CrossRef] [Green Version]
  249. Bruno, F.; Granata, V.; Bellisari, F.C.; Sgalambro, F.; Tommasino, E.; Palumbo, P.; Arrigoni, F.; Cozzi, D.; Grassi, F.; Brunese, M.C.; et al. Advanced Magnetic Resonance Imaging (MRI) Techniques: Technical Principles and Applications in Nanomedicine. Cancers 2022, 14, 1626. [Google Scholar] [CrossRef]
  250. Obaidat, I.M.; Narayanaswamy, V.; Alaabed, S.; Sambasivam, S.; Muralee Gopi, C.V.V. Principles of Magnetic Hyperthermia: A Focus on Using Multifunctional Hybrid Magnetic Nanoparticles. Magnetochemistry 2019, 5, 67. [Google Scholar] [CrossRef]
  251. Gavilán, H.; Simeonidis, K.; Myrovali, E.; Mazarío, E.; Chubykalo-Fesenko, O.; Chantrell, R.; Balcells, L.; Angelakeris, M.; Morales, M.P.; Serantes, D. How size, shape and assembly of magnetic nanoparticles give rise to different hyperthermia scenarios. Nanoscale 2021, 13, 15631–15646. [Google Scholar] [CrossRef]
  252. Xie, J.; Chen, K.; Huang, J.; Lee, S.; Wang, J.; Gao, J.; Li, X.; Chen, X. PET/NIRF/MRI triple functional iron oxide nanoparticles. Biomaterials 2010, 31, 3016–3022. [Google Scholar] [CrossRef] [Green Version]
  253. Wang, X.; Tu, M.; Tian, B.; Yi, Y.; Wei, Z.Z.; Wei, F. Synthesis of tumor-targeted folate conjugated fluorescent magnetic albumin nanoparticles for enhanced intracellular dual-modal imaging into human brain tumor cells. Anal. Biochem. 2016, 512, 8–17. [Google Scholar] [CrossRef]
  254. Pivetal, J.; Frénéa-Robin, M.; Haddour, N.; Vézy, C.; Zanini, L.F.; Ciuta, G.; Dempsey, N.M.; Dumas-Bouchiat, F.; Reyne, G.; Bégin-Colin, S.; et al. Development and applications of a DNA labeling method with magnetic nanoparticles to study the role of horizontal gene transfer events between bacteria in soil pollutant bioremediation processes. Environ. Sci. Pollut. Res. 2015, 22, 20322–20327. [Google Scholar] [CrossRef]
  255. Ali, R.S.; Meng, H.; Li, Z. Zinc-Based Metal-Organic Frameworks in Drug Delivery, Cell Imaging, and Sensing. Molecules 2022, 27, 100. [Google Scholar]
  256. Hossain, S.; Rahman, M.; Nahar, Y.; Rahman, A.; Sharafat, M.K.; Hossain, M.; Ochiai, B.; Elaissari, A.; Ahmad, H. A simple in situ synthesis of iron oxide magnetic nanoparticles embedded in thermosensitive polymer for DNA capture. J. Mater. Res. 2020, 35, 2441–2450. [Google Scholar] [CrossRef]
  257. Damavandi, F.; Wang, W.; Shen, W.Z.; Cetinel, S.; Jordan, T.; Jovel, J.; Montemagno, C.; Wong, G.K.S. Enrichment of low abundance DNA/RNA by oligonucleotide-clicked iron oxide nanoparticles. Sci. Rep. 2021, 11, 1–10. [Google Scholar] [CrossRef]
  258. Li, B.; Mou, X.; Chen, Z.; Chen, H.; Deng, Y.; Li, S.; Su, E.; He, L.; He, N. The development of a rapid high-quality universal nucleic acid extraction kit based on magnetic separation. Sci. China Chem. 2017, 60, 1602–1608. [Google Scholar] [CrossRef]
  259. Pinchon, E.; Leon, F.; Temurok, N.; Morvan, F.; Vasseur, J.J.; Clot, M.; Foulongne, V.; Cantaloube, J.F.; Perre, P.V.; Daynès, A.; et al. Rapid and specific DNA detection by magnetic field-enhanced agglutination assay. Talanta 2020, 219, 121344. [Google Scholar] [CrossRef] [PubMed]
  260. Camacho-Fernández, J.C.; Jin, M.; Liu, X.; Berg, A.V.D.; Zhou, G. Ultrasensitive DNA detection based on two-step quantitative amplification on magnetic nanoparticles. Nanotechnology 2016, 27, 335102. [Google Scholar]
  261. Yang, Z.; Shen, B.; Yue, L.; Miao, Y.; Hu, Y.; Ouyang, R. Application of Nanomaterials to Enhance Polymerase. Molecules 2022, 27, 8854. [Google Scholar] [CrossRef] [PubMed]
  262. Rocha-Santos, T.A.P. Sensors and biosensors based on magnetic nanoparticles. TrAC Trends Anal. Chem. 2014, 62, 28–36. [Google Scholar] [CrossRef]
  263. Sayad, A.; Skafidas, E.; Kwan, P. Magneto-impedance biosensor sensitivity: Effect and enhancement. Sensors 2020, 20, 5213. [Google Scholar] [CrossRef]
  264. Ramin, N.A.; Ramachandran, M.R.; Saleh, N.M.; Mat Ali, Z.M.; Asman, S. Magnetic Nanoparticles Molecularly Imprinted Polymers: A Review. Curr. Nanosci. 2022, 18, 1–29. [Google Scholar] [CrossRef]
  265. Meseguer-Lloret, S.; Torres-Cartas, S.; Gómez-Benito, C.; Herrero-Martínez, J.M. Magnetic molecularly imprinted polymer for the simultaneous selective extraction of phenoxy acid herbicides from environmental water samples. Talanta 2022, 239, 123082. [Google Scholar] [CrossRef]
  266. Fresco-Cala, B.; Batista, A.D.; Cárdenas, S. Molecularly imprinted polymer micro- And nano-particles: A review. Molecules 2020, 25, 4740. [Google Scholar] [CrossRef]
  267. Dmitrienko, E.V.; Pyshnaya, I.A.; Martyanov, O.N.; Pyshnyi, D.V. Molecularly imprinted polymers for biomedical and biotechnological applications. Russ. Chem. Rev. 2016, 85, 513–536. [Google Scholar] [CrossRef]
  268. Ariani, M.D.; Zuhrotun, A.; Manesiotis, P.; Hasanah, A.N. Magnetic Molecularly Imprinted Polymers: An Update on Their Use in the Separation of Active Compounds from Natural Products. Polymers 2022, 14, 1389. [Google Scholar] [CrossRef]
  269. Dmitrienko, E.V.; Bulushev, R.D.; Haupt, K.; Kosolobov, S.S.; Latyshev, A.V.; Pyshnaya, I.A.; Pyshnyi, D.V. A simple approach to prepare molecularly imprinted polymers from nylon-6. J. Mol. Recognit. 2013, 26, 368–375. [Google Scholar] [CrossRef]
  270. Dinc, M.; Basan, H.; Diemant, T.; Behm, R.J.; Lindén, M.; Mizaikoff, B. Inhibitor-assisted synthesis of silica-core microbeads with pepsin-imprinted nanoshells. J. Mater. Chem. B 2016, 4, 4462–4469. [Google Scholar] [CrossRef]
  271. Gao, S.; Wang, W.; Wang, B. Molecularly imprinted polymers as recognition elements in optical sensors. In Molecularly Imprinted Materials; CRC: Boca Raton, FL, USA, 2008; pp. 701–726. [Google Scholar] [CrossRef]
  272. Huynh, T.P.; Pieta, P.; D’Souza, F.; Kutner, W. Molecularly imprinted polymer for recognition of 5-fluorouracil by RNA-type nucleobase pairing. Anal. Chem. 2013, 85, 8304–8312. [Google Scholar] [CrossRef]
  273. Babamiri, B.; Salimi, A.; Hallaj, R. A molecularly imprinted electrochemiluminescence sensor for ultrasensitive HIV-1 gene detection using EuS nanocrystals as luminophore. Biosens. Bioelectron. 2018, 117, 332–339. [Google Scholar] [CrossRef]
  274. Slinchenko, O.; Rachkov, A.; Miyachi, H.; Ogiso, M.; Minoura, N. Imprinted polymer layer for recognizing double-stranded DNA. Biosens. Bioelectron. 2004, 20, 1091–1097. [Google Scholar] [CrossRef]
  275. Huang, L.; Wang, X.; Xie, X.; Xie, W.; Li, X.; Gong, X.; Long, S.; Guo, H.; Liu, Z. Synthesis and DNA Adsorption of Poly(2-Vinyl-4,6-Diamino-1,3,5-Triazine) Coated Polystyrene Microspheres. J. Wuhan Univ. Technol. Mater. Sci. Ed. 2018, 33, 999–1006. [Google Scholar] [CrossRef]
  276. Rutkowska, M.; Płotka-Wasylka, J.; Morrison, C.; Wieczorek, P.P.; Namieśnik, J.; Marć, M. Application of molecularly imprinted polymers in analytical chiral separations and analysis. Trends Anal. Chem. 2018, 102, 91–102. [Google Scholar] [CrossRef] [Green Version]
  277. Ding, X.; Heiden, P.A. Recent developments in molecularly imprinted nanoparticles by surface imprinting techniques. Macromol. Mater. Eng. 2014, 299, 268–282. [Google Scholar] [CrossRef]
  278. Whitcombe, M.J.; Kirsch, N.; Nicholls, I.A. Molecular imprinting science and technology: A survey of the literature for the years 2004–2011. J. Mol. Recognit. 2014, 27, 297–401. [Google Scholar] [CrossRef] [Green Version]
  279. Garnier, M.; Sabbah, M.; Ménager, C.; Griffete, N. Hybrid molecularly imprinted polymers: The future of nanomedicine? Nanomaterials 2021, 11, 3091. [Google Scholar] [CrossRef]
  280. Niu, M.; Pham-Huy, C.; He, H. Core-shell nanoparticles coated with molecularly imprinted polymers: A review. Microchim. Acta 2016, 183, 2677–2695. [Google Scholar] [CrossRef]
  281. Huang, S.; Xu, J.; Zheng, J.; Zhu, F.; Xie, L.; Ouyang, G. Synthesis and application of magnetic molecularly imprinted polymers in sample preparation. Anal. Bioanal. Chem. 2018, 410, 3991–4014. [Google Scholar] [CrossRef] [PubMed]
  282. Li, J.; Wang, Y.; Yu, X. Magnetic Molecularly Imprinted Polymers: Synthesis and Applications in the Selective Extraction of Antibiotics. Front. Chem. 2021, 9, 1–17. [Google Scholar] [CrossRef] [PubMed]
  283. Li, J.; Zhou, Q.; Yuan, Y.; Wu, Y. Iron-based magnetic molecular imprinted polymers and their application in removal and determination of di-n-pentyl phthalate in aqueous media. R. Soc. Open Sci. 2017, 4, 170672. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  284. Goyal, G.; Bhakta, S.; Mishra, P. Surface Molecularly Imprinted Biomimetic Magnetic Nanoparticles for Enantioseparation. ACS Appl. Nano Mater. 2019, 2, 6747–6756. [Google Scholar] [CrossRef]
  285. Dai, Q.; Wang, Y.; Xu, W.; Liu, Y.; Zhou, Y. Adsorption and specific recognition of DNA by using imprinted polymer layers grafted onto ionic liquid functionalized magnetic microspheres. Microchim. Acta 2017, 184, 4433–4441. [Google Scholar] [CrossRef]
  286. Sanadgol, N.; Wackerlig, J. Developments of smart drug-delivery systems based on magnetic molecularly imprinted polymers for targeted cancer therapy: A short review. Pharmaceutics 2020, 12, 831. [Google Scholar] [CrossRef]
  287. Nawaz, N.; Abu Bakar, N.K.; Mahmud, H.N.M.E.; Jamaludin, N.S. Molecularly imprinted polymers-based DNA biosensors. Anal. Biochem. 2021, 630, 114328. [Google Scholar] [CrossRef]
  288. Zhang, Z.; Liu, J. Molecularly Imprinted Polymers with DNA Aptamer Fragments as Macromonomers. ACS Appl. Mater. Interfaces 2016, 8, 6371–6378. [Google Scholar] [CrossRef]
  289. Brahmbhatt, H.; Poma, A.; Pendergraff, H.M.; Watts, J.K.; Turner, N.W. Improvement of DNA recognition through molecular imprinting: Hybrid oligomer imprinted polymeric nanoparticles (oligoMIP NPs). Biomater. Sci. 2016, 4, 281–287. [Google Scholar] [CrossRef] [Green Version]
  290. Zhang, Z.; Liu, J. Molecular Imprinting with Functional DNA. Small 2019, 15, 1805246. [Google Scholar] [CrossRef]
  291. Yazdanparast, S.; Benvidi, A.; Azimzadeh, M.; Tezerjani, M.D.; Ghaani, M.R. Experimental and theoretical study for miR-155 detection through resveratrol interaction with nucleic acids using magnetic core-shell nanoparticles. Microchim. Acta 2020, 187, 1–10. [Google Scholar] [CrossRef]
  292. Wang, L.; Liu, Z.J.; Cao, H.X.; Liang, G.X. Ultrasensitive colorimetric miRNA detection based on magnetic 3D DNA walker and unmodified AuNPs. Sens. Actuators B Chem. 2021, 337, 3–9. [Google Scholar] [CrossRef]
  293. Masud, M.K.; Umer, M.; Hossain, M.S.A.; Yamauchi, Y.; Nguyen, N.T.; Shiddiky, M.J.A. Nanoarchitecture Frameworks for Electrochemical miRNA Detection. Trends Biochem. Sci. 2019, 44, 433–452. [Google Scholar] [CrossRef]
  294. Li, P.; Li, M.; Zhang, F.; Wu, M.; Jiang, X.; Ye, B.; Zhao, Z.; Yue, D.; Fan, Q.; Chen, H. High-efficient nucleic acid separation from animal tissue samples via surface modified magnetic nanoparticles. Sep. Purif. Technol. 2021, 262, 118348. [Google Scholar] [CrossRef]
  295. Yang, Q.; Dong, Y.; Qiu, Y.; Yang, X.; Cao, H.; Wu, Y. Design of Functional Magnetic Nanocomposites for Bioseparation. Colloids Surf. B Biointerfaces 2020, 191, 111014. [Google Scholar] [CrossRef]
  296. Mdlovu, N.V.; Lin, K.S.; Chen, Y.; Wu, C.M. Formulation of magnetic nanocomposites for intracellular delivery of micro-RNA for MYCN inhibition in neuroblastoma. Colloids Surf. A Physicochem. Eng. Asp. 2021, 615, 126264. [Google Scholar] [CrossRef]
  297. Lawai, V.; Ngaini, Z. Chitosan magnetic nanocomposites for gene delivery. In Polysaccharide-Based Nanocomposites for Gene Delivery and Tissue Engineering; Woodhead: Sawston, UK, 2021; p. 335. [Google Scholar]
  298. Do, H.D.; Ménager, C.; Michel, A.; Seguin, J.; Korichi, T.; Dhotel, H.; Marie, C.; Doan, B.T.; Mignet, N. Development of theranostic cationic liposomes designed for image-guided delivery of nucleic acid. Pharmaceutics 2020, 12, 854. [Google Scholar] [CrossRef]
  299. Emaus, M.N.; Varona, M.; Eitzmann, D.R.; Hsieh, S.A.; Zeger, V.R.; Anderson, J.L. Nucleic acid extraction: Fundamentals of sample preparation methodologies, current advancements, and future endeavors. TrAC Trends Anal. Chem. 2020, 130, 115985. [Google Scholar] [CrossRef]
  300. Zheng, H.; Lin, H.; Chen, X.; Sui, J.; Ullah Khan, M.; Ramesh Pavase, T.; Han, X.; Cao, L. Tailor-made magnetic nanocomposite with pH and thermo-dual responsive copolymer brush for bacterial separation. Food Chem. 2021, 358, 129907. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of MNCs with different coatings, target ligands, drugs, etc. Molecules can be embedded in the coating or adsorbed on the surface.
Figure 1. Schematic illustration of MNCs with different coatings, target ligands, drugs, etc. Molecules can be embedded in the coating or adsorbed on the surface.
Magnetochemistry 09 00012 g001
Figure 2. MNC structure types. Multilayer shells and yolk–shell are the subtypes of core–shell MNCs. Assembly, Janus, and multicomponent are self-assembly structural types.
Figure 2. MNC structure types. Multilayer shells and yolk–shell are the subtypes of core–shell MNCs. Assembly, Janus, and multicomponent are self-assembly structural types.
Magnetochemistry 09 00012 g002
Figure 3. Schematic representation of polymer-coating features. Polymer shell usually provides higher colloidal stability and possible surface modification by functional, reporter, and targeting molecules (nucleic acids, antibodies, vitamins, etc.).
Figure 3. Schematic representation of polymer-coating features. Polymer shell usually provides higher colloidal stability and possible surface modification by functional, reporter, and targeting molecules (nucleic acids, antibodies, vitamins, etc.).
Magnetochemistry 09 00012 g003
Figure 4. Primary polymer-coated MNC structures.
Figure 4. Primary polymer-coated MNC structures.
Magnetochemistry 09 00012 g004
Figure 5. Synthesis of two types of silicaMNCs. Tetraethoxysilane (TEOS) forms a silicon dioxide shell. The simple hydrolysis of TEOS leads to a core–shell type structure. The presence of detergents, such as a cetrimonium bromide (CTAB), provides MNP encapsulation. (3-Aminopropyl)triethoxysilane (APTES) is a primary reagent for amino group surface functionalization.
Figure 5. Synthesis of two types of silicaMNCs. Tetraethoxysilane (TEOS) forms a silicon dioxide shell. The simple hydrolysis of TEOS leads to a core–shell type structure. The presence of detergents, such as a cetrimonium bromide (CTAB), provides MNP encapsulation. (3-Aminopropyl)triethoxysilane (APTES) is a primary reagent for amino group surface functionalization.
Magnetochemistry 09 00012 g005
Figure 6. Noble metals–MNC synthesis. MNCs can be obtained by mixing two types of nanomaterials or by spraying noble metal powder on MNPs. Forms such as core–shell, dumbbell, flower, octahedral, star, rod, and Janus may be formed.
Figure 6. Noble metals–MNC synthesis. MNCs can be obtained by mixing two types of nanomaterials or by spraying noble metal powder on MNPs. Forms such as core–shell, dumbbell, flower, octahedral, star, rod, and Janus may be formed.
Magnetochemistry 09 00012 g006
Figure 7. Carbon nanomaterial types that can be used for MNC production.
Figure 7. Carbon nanomaterial types that can be used for MNC production.
Magnetochemistry 09 00012 g007
Figure 8. Diagnostics, therapy, drug and gene delivery, biosensor, and magnetic separation applications of MNCs.
Figure 8. Diagnostics, therapy, drug and gene delivery, biosensor, and magnetic separation applications of MNCs.
Magnetochemistry 09 00012 g008
Figure 9. Possible toxicity mechanism of MNCs, leading cell processes’ dysregulation and triggering cell death (necrosis or apoptosis).
Figure 9. Possible toxicity mechanism of MNCs, leading cell processes’ dysregulation and triggering cell death (necrosis or apoptosis).
Magnetochemistry 09 00012 g009
Figure 10. Basic principles of NA magnetic separation.
Figure 10. Basic principles of NA magnetic separation.
Magnetochemistry 09 00012 g010
Figure 11. Basic principles of NA-based biosensors.
Figure 11. Basic principles of NA-based biosensors.
Magnetochemistry 09 00012 g011
Figure 12. Schematic representation of core–shell imprinting for MMIP preparation. The analyte is a template molecule that MIP has reversibly recognized. The number of cycles is proportional to the efficiency of MIPs.
Figure 12. Schematic representation of core–shell imprinting for MMIP preparation. The analyte is a template molecule that MIP has reversibly recognized. The number of cycles is proportional to the efficiency of MIPs.
Magnetochemistry 09 00012 g012
Table 1. Some examples of nucleic acid-based biomedical applications of MNCs.
Table 1. Some examples of nucleic acid-based biomedical applications of MNCs.
Application AreaMNCs TypeNA-Based ApplicationReference
Biosensing and diagnosticsMNP@Ag-amine-modified anti-miR-155miR-155 detection through resveratrol interaction (electrochemical label)[291]
MNP@AuUltrasensitive colorimetric and electrochemical miRNA detection[292,293]
MNP@grapheneElectrochemical miRNA detection[293]
MNP@SiO2DNA and RNA extraction from Hepatocellular Carcinoma, virus RNA extraction and detection by RT-PCR, Taq polymerase fixation for long-term enzyme activity for PCR[122,261]
MNP-oleic acidDNA detection by PCR[261]
MNP-NH2DNA extraction from blood and detection by PCR[122]
MNP-COOHDNA extraction from staphylococcus aureus bacteriophages, mRNA isolation from mammalian cells[122]
MNP-OH/-NH2/-COOHHybrid NA separation from animal tissue samples[294]
MMIPDNA detection[287]
MNP—rabbit antigoat immunoglobulinImmunoglobulin (IgG) detection[295]
Therapy and diagnosticsMNP@PEImicro-RNA intracellular delivery for MYCN inhibition in neuroblastoma[296]
MNP-chitosanGene delivery[297]
MNP-Hyaluronic acidGene delivery[298]
MNP-lipidssiRNA delivery[244]
MNP lipoplexTheranostics, imaging guided (MRI) delivery of NA[298]
Magnetic separationMNP@AgmRNA extraction[63]
MNP@AumRNA, dsDNA extraction[63,114]
MNP@graphenedsDNA extraction[114]
MNP@SiO2DNA/RNA extraction[62,63,114,122]
MNP@SiO2NA capture from lysed white blood cells, B. subtilis, E. coli, and Rift Valley fever viruses[299]
MNP@SiO2viral NA extraction from serum[122]
MNP@SiO2-organic halideDNA extraction[62]
MNP@SiO2-NH2DNA extraction[62]
MNP@polydopaminegenomic DNA extraction[171]
MNP-Nylon-6RNA extraction[19]
MNP-StreptavidinDNA/RNA extraction, aptamer-based cell separation[50,299]
MNP-CD138 (syndecan-1) antibody conjugatedEndothelial cells (HUVEC)
separation
[295]
MNP@PEIdsDNA extraction[114]
MNP-thermosensitive polymer, poly(N-isopropylacrylamide-co-2-aminoethyl methacrylate)DNA extraction[256]
MNP-N-isopropylacrylamide and allyl glycidyl ether, 3,5-difluoro-4-formylphenylboronic acidS. aureus and Salmonella spp. separation[300]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Popova, V.; Dmitrienko, E.; Chubarov, A. Magnetic Nanocomposites and Imprinted Polymers for Biomedical Applications of Nucleic Acids. Magnetochemistry 2023, 9, 12. https://doi.org/10.3390/magnetochemistry9010012

AMA Style

Popova V, Dmitrienko E, Chubarov A. Magnetic Nanocomposites and Imprinted Polymers for Biomedical Applications of Nucleic Acids. Magnetochemistry. 2023; 9(1):12. https://doi.org/10.3390/magnetochemistry9010012

Chicago/Turabian Style

Popova, Victoriya, Elena Dmitrienko, and Alexey Chubarov. 2023. "Magnetic Nanocomposites and Imprinted Polymers for Biomedical Applications of Nucleic Acids" Magnetochemistry 9, no. 1: 12. https://doi.org/10.3390/magnetochemistry9010012

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