Download as pdf or txt
Download as pdf or txt
You are on page 1of 29

processes

Review
Magnetic Iron Nanoparticles: Synthesis, Surface Enhancements,
and Biological Challenges
Jesús Roberto Vargas-Ortiz 1 , Carmen Gonzalez 2, * and Karen Esquivel 1, *

1 División de Investigación y Posgrado, Facultad de Ingeniería, Universidad Autónoma de Querétaro, Cerro de


las Campanas, Santiago de Queretaro 76010, Mexico
2 Facultad de Ciencias Quimicas, Universidad Autonoma de San Luis Potosi, Av. Dr. Manuel Nava Num. 6,
Zona Universitaria, San Luis Potosi 78210, Mexico
* Correspondence: gonzalez.castillocarmen@uaslp.mx (C.G.); karen.esquivel@uaq.mx (K.E.);
Tel.: +52-444-826-23-00 (ext. 6521) (C.G.); +52-442-192-12-00 (ext. 65401) (K.E.)

Abstract: This review focuses on the role of magnetic nanoparticles (MNPs), their physicochemical
properties, their potential applications, and their association with the consequent toxicological
effects in complex biologic systems. These MNPs have generated an accelerated development and
research movement in the last two decades. They are solving a large portion of problems in several
industries, including cosmetics, pharmaceuticals, diagnostics, water remediation, photoelectronics,
and information storage, to name a few. As a result, more MNPs are put into contact with biological
organisms, including humans, via interacting with their cellular structures. This situation will
require a deeper understanding of these particles’ full impact in interacting with complex biological
systems, and even though extensive studies have been carried out on different biological systems
discussing toxicology aspects of MNP systems used in biomedical applications, they give mixed and
inconclusive results. Chemical agencies, such as the Registration, Evaluation, Authorization, and
Restriction of Chemical substances (REACH) legislation for registration, evaluation, and authorization
Citation: Vargas-Ortiz, J.R.; Gonzalez,
of substances and materials from the European Chemical Agency (ECHA), have held meetings to
C.; Esquivel, K. Magnetic Iron discuss the issue. However, nanomaterials (NMs) are being categorized by composition alone,
Nanoparticles: Synthesis, Surface ignoring the physicochemical properties and possible risks that their size, stability, crystallinity, and
Enhancements, and Biological morphology could bring to health. Although several initiatives are being discussed around the
Challenges. Processes 2022, 10, 2282. world for the correct management and disposal of these materials, thanks to the extensive work of
https://doi.org/10.3390/pr10112282 researchers everywhere addressing the issue of related biological impacts and concerns, and a new
Academic Editors: Huseyin
nanoethics and nanosafety branch to help clarify and bring together information about the impact of
Tombuloglu, Abdulhadi Baykal nanoparticles, more questions than answers have arisen regarding the behavior of MNPs with a wide
and Fabio Carniato range of effects in the same tissue. The generation of a consolidative framework of these biological
behaviors is necessary to allow future applications to be manageable.
Received: 4 October 2022
Accepted: 2 November 2022
Keywords: iron oxide; magnetic nanoparticles; nanomaterials; nanoparticles; nanotoxicology
Published: 4 November 2022

Publisher’s Note: MDPI stays neutral


with regard to jurisdictional claims in
published maps and institutional affil- 1. Introduction
iations.
Nanoparticles (NPs) and nanostructured materials have innovative properties due
to their small size, large surface area, crystalline arrangement, composition, and reac-
tivity [1,2]. These properties confer unique and specific effects, such as superparamag-
Copyright: © 2022 by the authors.
netism, surface area and charge, and optic and magnetic properties different from their
Licensee MDPI, Basel, Switzerland. bulk presentation [3–7]. The design of these materials influences the final properties they
This article is an open access article present. Due to the different crystalline planes exposed on the surface of the said particles,
distributed under the terms and they could externalize other atoms of the structure, which interact differently with the
conditions of the Creative Commons environment where they are found [8,9].
Attribution (CC BY) license (https:// A well-known example is magnetic materials; when found on the nanoscale, iron
creativecommons.org/licenses/by/ oxide nanoparticles (IONPs) obtain a single-domain phenomenon known as superpara-
4.0/). magnetism [10,11] that can shorten the relaxation times as contrast agents in magnetic

Processes 2022, 10, 2282. https://doi.org/10.3390/pr10112282 https://www.mdpi.com/journal/processes


Processes 2022, 10, 2282 2 of 29

resonance imaging (MRI) [12,13] travel to target locations with magnetic fields [11] and
simultaneously play a role in both diagnostics and treatment, known as theranostics, by
vibrating and heating. Recent design strategies of IONP-based nanoplatforms describe
the rationale for the combination of other functional materials with IONPs and test their
possible applications in smart nanomedicine theranostics [14]. However, dispersed NPs
tend to form unstable suspensions in culture media, and modifying their surface chemistry,
for example, with rhodium citrate, can increase their adverse effects, cytotoxicity, and
uptake in breast cancer cell cultures [15]. Whether particles can become toxic by increasing
their size (agglomerates) or by maintaining their nanometric size in dispersions is debated.
Nevertheless, the increasing size implies a reduction in the number of atoms on the surface;
thus, they become bulk materials and have reduced harmful effects [16–22].
Size is crucial when designing structures that interact at some point with the human
body [23–25]. The range of cell sizes in the human body is wide, from the granules of the
cerebellum in the brain (5 µm) up to the ovule (0.1 mm). Between these cells, a great variety
of sizes are among the different systems responsible for maintaining homeostasis within
the body. All these cells are three orders of magnitude larger than an NP [26]. Considering
that one main application of magnetic nanoparticles is in diagnostics and treatment, such
as imaging, the importance of these interactions becomes paramount to consent.
Imaging plays a crucial role in diagnosis because of the importance of the swiftness,
efficiency, and accuracy of reliable early detection of diseases within the human body,
such as cancer and inflammation [27]. Various contrast agents have been used to better
visualize resolution by increasing the contrast between healthy and abnormal tissue that
occurs within the human body, such as gadolinium, manganese, and barium [28]. Con-
trast agents for MRI are classified according to their specific characteristics, including:
(i) chemical composition, (ii) route of administration, (iii) magnetic properties, (iv) effect
on the magnetic resonance imaging, (v) biodistribution, and (vi) imaging applications
(target/organ-specific agents) [29]. MNPs provide a potential promising alternative due to
their improved magnetic characteristics (superparamagnetism) in terms of specificity, ease
of functionalization, and biocompatibility.
Most of these agents are complexes of lanthanide elements such as Gd3+ or transition-
metal manganese (Mn2+ ) with superparamagnetic magnetite particles [30]. These factors
are used to shorten T1 or T2 relaxation times; the signal intensity on T1-weighted images
is enhanced, or the signal intensity on T2-weighted images is decreased [31]. T1 and
T2 are two different relaxation durations used to describe abnormalities in tissue. Most
paramagnetic contrast agents are positive agents, which shorten the T1 relaxation time so
that enhancements appear bright on T1-weighted images. However, these agents are used
in an average of one-third of the analyses, given their low precision, problems brought on by
the release of free gadolinium (Gd3+ ) [29], and the risk of presenting severe adverse effects in
pediatric patients [32]. Dysprosium, superparamagnetic agents, and ferromagnetic agents
are negative contrast agents [33,34], while enhancing parts appear darker on T2-weighted
images [31]. MRI contrast agents incorporate chelating agents to reduce storage in the
human body, improve excretion, and reduce adverse effects [35]. However, they must be
used immediately due to their short shelf life [36].
Magnetite and maghemite are nowadays regarded as safe. They are now in clinical
usage as MRI contrast agents [37–39], pharmaceutical agents [40,41], and a combination of
both treatment and diagnostics, also known as theranostics [42–44], due to their significant
superparamagnetic characteristics (around 40 to 90 Am2 /kg) achievable because of their
nanometric size (10nm) (Figure 1).
Magnetite and maghemite are the most suitable MNPs, but their iron oxidation states
and structures differ, affecting their final physicochemical characteristics [45]. Other types
of magnetic nanoparticles exist with similar promising properties, including titanomag-
netite cobalt and nickel-based magnetic materials [46,47]. Nevertheless, the full impact
of the nanoparticles in general and in the case of magnetic materials is still under study
to correlate MNPs’ size, morphology, concentration, and crystallinity with different bio-
Processes 2022, 10, 2282 3 of 29

logical specimens. A collection of research from different cell lines and tissues has shown
inconsistent results from cell–MNP interactions [48], such as uptake by mononuclear cell
infiltration in mice liver [49–52], leaching of different ions such as Fe2+ [53–55], apoptosis
Processes 2022, 10, x FOR PEER REVIEW  3  of  29 
  via over-endocytosis, [56–58], oxidative stress by passive diffusion, receptor-mediated
endocytosis unchaining mitochondrial stress [59,60], and DNA damage leaching Fe2+ into
the cell’s core, increasing temperature while vibrating [61–63].

 
Figure 1. MNPs’ biomedical applications.
Figure 1. MNPs’ biomedical applications. 
This array of properties that NMs possess have made it possible to generate significant
advances in different areas, specifically biomedicine. For example, in the treatment and
Magnetite and maghemite are the most suitable MNPs, but their iron oxidation states 
diagnosis of diseases, the role of MNPs has become increasingly important in the last
and structures differ, affecting their final physicochemical characteristics [45]. Other types 
two decades, but the short-, medium-, and long-term toxicological implications are set
of magnetic nanoparticles exist with similar promising properties, including titanomag‐
aside. This review focuses on the collection of various studies on assessment and challenges
netite cobalt and nickel‐based magnetic materials [46,47]. Nevertheless, the full impact of 
with regard to synthesis routes, biological evaluation, and the impact of iron MNPs to
the nanoparticles in general and in the case of magnetic materials is still under study to 
gather their different characteristics and properties, and studies presenting the possible
correlate MNPs’ size, morphology, concentration, and crystallinity with different biolog‐
effects in various
ical  specimens.  A biological
collection assays and assessments
of  research  to raise
from  different  cell awareness
lines  and of the correct
tissues  use and
has  shown 
disposal of MNPs to generate an integrative background of knowledge enabling improved
inconsistent results from cell–MNP interactions [48], such as uptake by mononuclear cell 
visualization of the context behind the biological behavior of the iron
infiltration in mice liver [49–52], leaching of different ions such as Fe MNPs.
2+ [53–55], apoptosis 

via over‐endocytosis, [56–58], oxidative stress by passive diffusion, receptor‐mediated en‐
2. Magnetic Nanoparticle Synthesis Methods
docytosis unchaining mitochondrial stress [59,60], and DNA damage leaching Fe2+ into 
For manufacturing quality control and achieving the required outcomes for different
the cell’s core, increasing temperature while vibrating [61–63]. 
applications, defined parameters for NPs’ diverse characteristics are essential. These NPs,
This array of properties that NMs possess have made it possible to generate signifi‐
composed of magnetic iron oxide, Fe3 O4 (magnetite), and γ-Fe2 O3 (maghemite), among
cant advances in different areas, specifically biomedicine. For example, in the treatment 
other common structures with modified surface chemistry, are considered promising
and diagnosis of diseases, the role of MNPs has become increasingly important in the last 
materials
two  for but 
decades,  widespread industrial
the  short‐,  applications
medium‐,  in computer
and  long‐term  and sensing
toxicological  materialsare 
implications  [64–66],
set 
the textile industry [67,68], ferrofluids [69,70], the remediation industry for the
aside. This review focuses on the collection of various studies on assessment and chal‐ degradation
of herbicides and emergent contaminants [71–73] and intrinsic magnetic properties [74,75],
lenges with regard to synthesis routes, biological evaluation, and the impact of iron MNPs 
the biomedical area for iron deficiency anemia [76,77], imaging and diagnostics [78–81],
to gather their different characteristics and properties, and studies presenting the possible 
heat cauterization treatments [82–84], and drug delivery [81,85–88] due to their chemical
effects in various biological assays and assessments to raise awareness of the correct use 
nature and biocompatibility. The unique characteristics that MNPs possess depend on the
and disposal of MNPs to generate an integrative background of knowledge enabling im‐
synthesis conditions.
proved visualization of the context behind the biological behavior of the iron MNPs. 
Therefore, it is crucial to examine nanomaterials based on the physicochemical charac-
teristics they acquired from the synthesis route used to obtain them (Figure 2).
2. Magnetic Nanoparticle Synthesis Methods 
Like most nanoparticles, MNPs have a high surface-area-to-volume ratio, implying
theyFor manufacturing quality control and achieving the required outcomes for different 
have qualities that are different from those of bulk material, such as a lower melting
applications, defined parameters for NPs’ diverse characteristics are essential. These NPs, 
point, a lower sintering temperature, and distinct magnetic properties [35,89–91].
composed  of magnetic iron 
Different MNPs can beoxide,  Fe3by
obtained O4 acquiring
(magnetite),  and morphologies
unique γ‐Fe2O3  (maghemite), 
depending among 
on the
other common structures with modified surface chemistry, are considered promising ma‐
synthesis route and their conditions, such as temperature, pressure, reaction time, reagent
terials for widespread industrial applications in computer and sensing materials [64–66], 
casting, etc. (Figure 3), that could modify the MNPs’ physicochemical properties, surface
the textile industry [67,68], ferrofluids [69,70], the remediation industry for the degrada‐
functionalization, and magnetic properties. A ferromagnetic nanoparticle’s behavior is
tion of herbicides and emergent contaminants [71–73] and intrinsic magnetic properties 
highly influenced by its size. When the size is reduced, it transforms from a multidomain
[74,75], the biomedical area for iron deficiency anemia [76,77], imaging and diagnostics 
particle to a single-magnetic-domain nanoparticle and finally to a superparamagnetic
[78–81], heat cauterization treatments [82–84], and drug delivery [81,85–88] due to their 
chemical nature and biocompatibility. The unique characteristics that MNPs possess de‐
pend on the synthesis conditions.   
Therefore, it is crucial to examine nanomaterials based on the physicochemical char‐
acteristics they acquired from the synthesis route used to obtain them (Figure 2). 
Processes 2022, 10, 2282 4 of 29

Processes 2022, 10, x FOR PEER REVIEW  4  of  29
  nanoparticle [81]. All the latter properties are achievable through different methods of
synthesis to control the shape and coating in the case of functionalized materials.

Figure 2. Graphical representation of the available synthesis routes used to produce M

Like most nanoparticles, MNPs have a high surface‐area‐to‐volume ratio
they have qualities that are different from those of bulk material, such as a low
point, a lower sintering temperature, and distinct magnetic properties [35,89–9
Different MNPs can be obtained by acquiring unique morphologies dep
the synthesis route and their conditions, such as temperature, pressure, reactio
agent casting, etc. (Figure 3), that could modify the MNPs’ physicochemical 
surface functionalization, and magnetic properties. A ferromagnetic nanopartic
ior is highly influenced by its size. When the size is reduced, it transforms fr
tidomain  particle to a  single‐magnetic‐domain  nanoparticle and finally to a 
magnetic nanoparticle [81]. All the latter properties are achievable through diffe
 
ods of synthesis to control the shape and coating in the case of functionalized 
Figure 2. Graphical representation of the available synthesis routes used to produce MNPs.
Figure 2. Graphical representation of the available synthesis routes used to produce MNPs. 

Like most nanoparticles, MNPs have a high surface‐area‐to‐volume ratio, implying
they have qualities that are different from those of bulk material, such as a lower melting
point, a lower sintering temperature, and distinct magnetic properties [35,89–91]. 
Different MNPs can be obtained by acquiring unique morphologies depending on
the synthesis route and their conditions, such as temperature, pressure, reaction time, re‐
agent casting, etc. (Figure 3), that could modify the MNPs’ physicochemical properties,
surface functionalization, and magnetic properties. A ferromagnetic nanoparticle’s behav‐
ior is highly influenced by its size. When the size is reduced, it transforms from a mul‐
tidomain  particle to a  single‐magnetic‐domain  nanoparticle and finally to a superpara‐
magnetic nanoparticle [81]. All the latter properties are achievable through different meth‐
ods of synthesis to control the shape and coating in the case of functionalized materials. 

 
Figure 3. Different shapes can be obtained when synthesizing MNPs, changing their
Figure 3. Different shapes can be obtained when synthesizing MNPs, changing their p
physicochemical properties.
ical properties. 
Several synthesis routes have been created to obtain different morphologies, crystal
growth, sizes, magnetic saturation, and surface charges of MNPs using parameters such as
Several synthesis routes have been created to obtain different morpholog
temperature, pressure, sonochemistry, and biological assistance with equipment, which
growth, sizes, magnetic saturation, and surface charges of MNPs using param
can be simple to acquire, such as those obtained via an ultrasonic bath, to relatively
as  temperature, 
complex pressure, 
microwave devices. sonochemistry, 
These routes and MNP
can provide specific biological 
propertiesassistance 
for different with  e
applications (Table 1).

   
Figure 3. Different shapes can be obtained when synthesizing MNPs, changing their physicochem‐
ical properties. 
Processes 2022, 10, 2282 5 of 29

Table 1. Comparison between synthesis routes for the formation of MNPs *.

Synthesis Temperature Size Shape Efficiency Magnetite


Environment Time Ref.
Route (◦ C) Control Control Output (XRD Pattern)
Insert at- Relatively Precursor-
Coprecipitation <100 Minutes Bad High [92–94]
mosphere broad dependent
Thermal
Insert at- Hours to Oxygen-
Aqueous routes

decomposi- 100–300 Excellent Excellent High [95–97]


mosphere days dependent
tion
High Hours to Temperature-
Hydrothermal 150–200 Excellent Excellent High [98]
pressure days dependent
Poor magnetite
Sol–gel 100–300 Ambient Hours Good Good Medium [99–102]
presence
Poor magnetite
Microemulsion <100 Ambient Hours Good Good Low [103–105]
presence
Assisted routes

Cavitation- and
Sonochemical
<50 Ambient Minutes Good Bad Medium frequency- [106,107]
assisted
dependent
Microwaved Minutes to High magnetite
100–200 Ambient Medium Good Medium [108,109]
assisted hours presence
Biologic-
Bacteria Room Hours to
Ambient Broad Bad Low assistant- [110–112]
Biologic routes

driven temp. days


dependent
Medium
magnetite
Room Relatively
Green Ambient Minutes Good Low presence. [113–115]
temp. good
Leaves nature-
dependent
* Table modified from [116].

2.1. Wet-Chemical Methods


2.1.1. Coprecipitation Synthesis
Coprecipitation is the ideal method for synthesis simplicity. This synthesis is a
straightforward, well-known, cheap, and effortless process for obtaining MNPs (Fe3 O4 and
γ-Fe2 O3 ) from a usual 1:2 molar ratio of ferric/ferrous salt solution by adding basic-pH
solutions at standard temperature or increased temperature [117,118]. Most of the reac-
tions could be helped with an inert gas atmosphere to prevent additional oxidation and
to maintain their magnetic properties [119,120]. The ferric/ferrous ratio, the salt utilized,
such as chlorides, sulfates, or nitrates, the reaction temperature, the pH value, and the
ionic strength of the medium all influence the size, morphology, and composition of the
magnetic particles [117], making the quality of magnetic particles entirely repeatable if the
synthesis conditions are controlled. On the other hand, control over particle size attributes
is reduced due to the rapid creation of particles. Overall, the coprecipitation methodology
creates abundant MNPs, with strong magnetic properties (50 to 90 Am2 kg−1 ) but poor
control over shape and size compared to other synthesis methods.

2.1.2. Thermal Decomposition


Thermal decomposition of organo-metallic precursor compounds in different solvents
(organic and water) with high boiling points and stabilizing surfactants is a potential
strategy for producing monodispersed magnetite nanoparticles without particle aggrega-
tion [83,121,122].
Thanks to the nucleation process, several synthesis variants have been developed for
synthesizing MNPs with controllable size and morphology thanks to the nucleation process.
Additionally, by promoting the adsorption of additives, e.g., the oleic acid–oleylamine
ligand pair, the growth of nanocrystals can be achieved, inspired by the synthesis of high-
quality semiconductor nanocrystals and oxides in a wet-chemical medium through thermal
decomposition [95,97,121,123]. As a result, smaller superparamagnetic nanoparticles can be
Processes 2022, 10, 2282 6 of 29

obtained. The nucleation condition of this method provides the possibility of the particles
being separated from growth, unlike coprecipitation, preventing a more complex reaction
process that can lead to oxidation [121]. In broad terms, the thermal decomposition synthe-
sis process is one of the most appraised in terms of adequate size and morphological control.
Thermal decomposition allows control over particle size, strong magnetic properties (40 to
90 Am2 kg−1 ), and high yield with easy access to precursors, with oxygen concentration
being the crucial factor in producing the desired iron oxide structure.

2.1.3. Hydrothermal Synthesis


One of the most frequent ways of preparing NMs is hydrothermal synthesis. Straight-
forwardly, hydrothermal synthesis synthesizes single-crystal nanoparticles relying on the
solubility of an aqueous solution (water) in the traditional crystal-making synthesis involv-
ing high temperatures and pressures [124]. It is essentially a solution-reaction-based synthe-
sis method. The creation of NMs in hydrothermal synthesis can occur over a wide range of
temperatures, from ambient temperature to extremely high temperatures (130–250 ◦ C). The
different vapor pressures in the reaction, in either low-pressure or high-pressure conditions
(0 and 3–4 MPa), can be utilized to control the morphology of the materials to be synthe-
sized, crystallizing the material in a sealed container [125,126] and highly improving the
preferential growth of different planes of the MNPs [127]. Nevertheless, the high pressure
and temperatures sacrifice the magnetic properties of the MNPs compared to other, cheaper,
and simpler synthesis routes.

2.1.4. Sol–Gel Synthesis


In the materials science and engineering field, sol–gel synthesis is a traditional and
commonly used technique in the wet-chemical approach, mainly used for NM fabrication
(typically metal oxides) [101,102]. It usually begins with a colloidal solution that serves
as a precursor for an integrated network of isolated particles. The system is shaped by a
sol phase, a stable dispersion of colloidal particles in a solvent, and a gel, a continuous
three-dimensional network that encloses the colloidal liquid phase. The aggregation of
colloidal particles forms the network in a colloidal gel. The particles in a polymer gel have
a polymeric sub-structure formed by the accumulation of sub-colloidal aggregates [116].
There are definite advantages of the sol–gel reaction over the previous wet-chemical or
aqueous synthesis method, such as the simple dispersion in solvents, due to hydrophilic
ligands on the surface of MNPs, which are used in the surface coating and for high control
over size and shape; however, the high costs of the reagent precursors (i.e., alkoxides), the
low yield compared to other techniques, such as coprecipitation, and poor obtention of
magnetite structure limit the access of the sol–gel synthesis.

2.1.5. Microemulsion Synthesis


The self-assembly synthesis methodology of micro- and NMs has piqued interest in
recent decades because of its unique features compared to bulk material analogs due to
the mechanical and physicochemical property changes at the micro- and nanoscale. A
surfactant stabilizes micro-domains in a thermodynamically stable and isotropic dispersion
of two immiscible liquids (usually water and an oil-based compound) (Figure 4). It includes
various structures and combinations that correlate to various self-assembled colloidal
system performances [128]. The vital interaction to consider is the surfactant stabilizing
the formed domains, which should be carefully chosen for the desired material. Iron oxide
nanoparticles have been synthesized through this method. Iron ionic precursors, such as
chlorides, in a basic-pH medium, such as ammonium, can be dissolved in water to create a
reaction-ready space capable of producing the desired particle condensation [104,129,130].
This method offers control of the size of the nanoparticle through the solution of precursors
and the available space provided in these solutions by adjusting the water-to-surfactant
ratio, which entails a chemical reaction that converts soluble precursors to an insoluble
in the aqueous phase, thus altering nanoparticle size [128,130]. Although monodispersed
Processes 2022, 10, x FOR PEER REVIEW  7  of  29 
 

Processes 2022, 10, 2282 7 of 29


water‐to‐surfactant ratio, which entails a chemical reaction that converts soluble precur‐
sors to an insoluble in the aqueous phase, thus altering nanoparticle size [128,130]. Alt‐
hough monodispersed nanoparticles can be synthesized through microemulsion with rea‐
nanoparticles can be synthesized through microemulsion with reasonable control over size
sonable control over size and shape, this technique has the lowest efficiency output, poor 
and shape, this technique has the lowest efficiency output, poor magnetite presence,
magnetite presence, and, consequently, low magnetic properties (20 to 60 Am 2kg−1).  and,
consequently, low magnetic properties (20 to 60 Am2 kg−1 ).

 
Figure 4. Representative micro-domains of the two immersible phases in microemulsions.
Figure 4. Representative micro‐domains of the two immersible phases in microemulsions. 

2.2. Assisted Methods


2.2. Assisted Methods 
2.2.1. Sonochemically Assisted
2.2.1. Sonochemically Assisted 
Acoustic cavitation is defined as the creation, expansion, and implosive collapse of
Acoustic cavitation is defined as the creation, expansion, and implosive collapse of 
bubbles in liquids bombarded with high-intensity sound, resulting in a transitory and
bubbles in liquids bombarded with high‐intensity sound, resulting in a transitory and lo‐
localized temperature of 5000 K and a pressure of 1000 bar [85,131]. This collapse results
calized temperature of 5000 K and a pressure of 1000 bar [85,131]. This collapse results in 
in intense localized heat and pressures with minimal periods [132], commonly known as
intense 
hotspots.localized  heat  and  pressures 
This technology produces with  minimal 
structures usingperiods  [132],  commonly 
the chemical known  as It
effect of ultrasound.
hotspots. This technology produces structures using the chemical effect of ultrasound. It 
provides a novel route without the use of high temperatures, high pressures, or extended
provides a novel route without the use of high temperatures, high pressures, or extended 
reaction periods, ideal for the formation of materials with complex structures sensitive to
reaction periods, ideal for the formation of materials with complex structures sensitive to 
extreme heat conditions (magnetic materials) [133,134]. The ultrasonic methodology route
extreme heat conditions (magnetic materials) [133,134]. The ultrasonic methodology route 
depends strongly on the cavitation and amplitude parameters applied to the particles. It
depends strongly on the cavitation and amplitude parameters applied to the particles. It 
can be observed sometimes as a complementary step in other synthesis routes to improve
can be observed sometimes as a complementary step in other synthesis routes to improve 
specific properties, such as coprecipitation, and to provide rapid reaction times or con-
specific properties, such as coprecipitation, and to provide rapid reaction times or control 
trol of emulsion factors [132,133,135], producing the medium output of strong magnetic
of emulsion factors [132,133,135], producing the medium output of strong magnetic na‐
nanoparticles without the shape and size control of other techniques.
noparticles without the shape and size control of other techniques. 
2.2.2. Microwave-Assisted
2.2.2.   Microwave‐Assisted 
The orientation of molecules that occur when irradiated with any source of electro-
magnetic waves, aligning their dipoles within the external field, creates a strong internal
The orientation of molecules that occur when irradiated with any source of electro‐
heat resulting from the excitation with electromagnetic radiation [136,137]. The shape
magnetic waves, aligning their dipoles within the external field, creates a strong internal 
and size of the obtained nanoparticles, as well as their magnetic characteristics, may vary
heat resulting from the excitation with electromagnetic radiation [136,137]. The shape and 
slightly
size  depending
of  the  on the experimental
obtained  nanoparticles,  as  well parameters, such ascharacteristics, 
as  their  magnetic  the precursor may material or
vary 
reaction temperature [109,136,138–141]. Microwave heating of a solution of iron oxide
slightly depending on the experimental parameters, such as the precursor material or re‐
nanostructures
action  such
temperature  as magnetite andMicrowave 
[109,136,138–141].  hematite hasheating 
been usedof  a tosolution 
improveof  their magnetic
iron  oxide 
properties, sacrificing crystal structure [138,142]. This process has a variety
nanostructures such as magnetite and hematite has been used to improve their magnetic  of advantages
over conventional heating-based procedures, such as high reaction rates and excellent prod-
properties, sacrificing crystal structure [138,142]. This process has a variety of advantages 
uct yields,
over  due to the
conventional  rapid localized
heating‐based  heating provided
procedures,  by microwaves
such  as  high  [143,144],
reaction  rates  reducing
and  excellent 
process time and energy costs compared to the conventional thermal methods, making
product yields, due to the rapid localized heating provided by microwaves [143,144], re‐
microwave-assisted
ducing  synthesis
process  time  and  energy methods more appealing
costs  compared  to  the  for iron oxide thermal 
conventional  nanoparticles.
methods, 
Both approaches are relatively simple processes in time and reaction times, obtaining
specific control over the nanoparticles.

 
Processes 2022, 10, 2282 8 of 29

2.3. Biological Synthesis Routes


Several researchers are now looking at creating dependable, clean, and environmen-
tally acceptable experimental procedures for NP synthesis. The biosynthesis of NPs uti-
lizing various biological components, such as plant derivatives, is a strategy with a lot
of promise.
Oxidation and other thermic methods tend to sacrifice structure to gain control over
the synthesis, size, magnetic properties, morphology, etc. [18,88,141,145]. Several types
of magnetotactic bacteria can create biologically manufactured domain-specific iron ox-
ide nanoparticles called magnetosomes via biomineralization, which can be caused by
Fe(III)-reducing bacteria such Thermoanaerobacter, Bdellovibrionota, Proteobacteria (alpha,
eta, zeta, gamma), SAR324, and Phoenix dactylifera species [146–149], typically grown
from a ferric oxyhydroxide precursor at 65 ◦ C. The resulting magnetite crystals, 5–140 nm
in size, are confined in a biological membrane regulating the magnetosome’s size and
form [148,150,151]. Magnetosomes in bacteria are grouped in chains with the [111] crystal
plane with narrow down-size distribution [152].
Green synthesis of NMs can result from, for instance, bacteria-driven synthesis of
living organisms’ capacity to reduce metal and oxides into defined structures within their
metabolism. Additionally, different plant species have biological extracts or the capability
to hyper-accumulate metal ions in tissues to form nanoparticles or decrease them in various
organs to generate metal oxide structures, as shown through different phytoremediation
studies [153]. As a result, many extracts from other plant parts, such as the leaves, stem, root,
fruit, and seed, have been employed to synthesize nanoparticles. Using natural extracts
of plant parts is a very cost-effective and environmentally friendly approach. It does not
require the use of any intermediate base groups such as silane (Si-H), mercaptoundecanoic
acid (MUA), oleic acid, cetyltrimethylammonium bromide (CTAB), polyvinylpyrrolidone
(PVP) [154], or sodium cholate, among others [155]. Likewise, expensive equipment and
precursors are not required. Because of the variety of metabolites in the plant extracts, such
as phenolic compounds that have ion-reducing capabilities, most of the efforts associated
with obtaining MNPs using plant extracts have been through the implementation of the
leaves as the source of these extracts [156–159]. Several aspects must be considered when
producing NMs through leaf extractions: the solvent, the plant origin, extraction methods,
reaction time, and external industrial chemical reagents [155,160].
MNPs are typically functionalized with various substances (such as citrate, phosphate,
chitosan, polyethylene glycol, albumin, silica, iron oxides, and metals) to increase colloidal
stability and biocompatibility since uncoated MNPs have the propensity to flocculate when
exposed to physiological fluids [161] and tend to present unfavorable effects [162]. The
assessment of the size and zeta potential as a function of time in various media (culture
medium, phosphate buffer, in the absence or presence of fetal bovine serum (FBS)) will help
us to understand how MNPs behave when they are in contact with bio-relevant culture
media. Although it is a novel approach to NP production, excessive energy consumption,
long reaction times, and the difficulty of obtaining precise chemical reactions to describe the
mechanisms underlying the biosynthesis process are the primary issues with the synthesis
process [163,164].

2.4. Surface Coating


Just like bulk materials, iron materials are prone to rust and corrosion. Iron oxide
nanoparticles cannot be used as a core material to produce treatments or pharmacological
agents per se without the addition of functional [82,165], biocompatible [166], inert, or
target coatings [167,168]. MNP design is essential in biomedical applications. The features
of the MNP core are determined by form and size, but the choice of added ligands and
coatings significantly impacts the colloidal stability and functions [169], further specifying
their properties for later applications.
It is a known fact that MNPs have an inherent instability when preserved over long
periods due to their magnetism [170–172], which leads to a loss in dispersibility because of
Processes 2022, 10, 2282 9 of 29

the aggregation and formation of large particles and reduces the surface energy, gen-
erating a loss of magnetism. Non-coated iron nanoparticles easily oxidize when are
exposed to air due to their high chemical activity and unstable spinel structure, espe-
cially magnetite and maghemite nanoparticles. They have a maximum shelf life of about
8 to 12 months [116,173]. As a result, chemically stabilizing simple iron oxide nanoparticles
to prevent damage during or after their subsequent exposition and/or application becomes
an essential factor to consider because most biological organisms and applications require
aqueous solutions to be administered [44,174]. Generally speaking, polymers, SiO2 , carbon-
derived structures, and metal material coats of a nanoscale substance constitute a family of
composite nanoparticles [175] to reduce their negative impacts on living organisms and
serve as ligands for functionalization [167].

2.4.1. Silica Coating


Due to its chemical stability, diversity in surface alterations, and limited loss of ini-
tial magnetism, silica (SiO2 ) has been regarded as the best material to safeguard MNPs.
Silica is beneficial in attaching different biological ligands to the iron oxide surface. Silica
coatings, metal, or metal oxide coatings help stabilize the nanoparticles in the solution.
Silica coating is the preferred material when applying layers to MNPs due to its increased
biocompatibility, ability to bond with organic linkers, and simple synthesis methodol-
ogy [176,177]. Different methodologies for coating MNPs and nanoparticles are based on
the desired properties, thickness, silica percentage, and later functionalization [100]. The
Stöber methodology is the most used and straightforward because of its simple application.
It requires a sol–gel precursor, which is hydrolyzed and condensed to obtain SiO2 sheets
that will cover the desired material [178]. A modified microemulsion method using a multi-
layer silica coating applied to nanoparticles with relative simplicity, assisted by microwave
radiation, has proven very effective at controlling single layers of a few atoms [179].

2.4.2. Carbon-Based Coatings


By adding functional groups to the iron oxide surface, such as hydroxyl (-OH), carboxyl
(-COOH), amino (-NH2 ), or silane (-SH), various bioactive compounds can be bonded. Due
to the benefits of linking several kinds of molecules to the coating [174,180], carbon coatings
are popular. These provide a magnetic core material with strong chemical and thermal
stability, high intrinsic electric conductivity, and an efficient oxidation barrier. Hydrophilic
carbon coatings also provide enhanced stability and dispersibility. With the co-precipitation
and surface oxidation of CNTs, nanoparticles have been effectively created, resulting in
hybrid materials with improved conductive properties from 16 meV at 20 ◦ C to a maximum
of 67 meV at 120 ◦ C [181]. Different synthesis methods can be used to produce core–shell
nanostructures. The three-step process uses MNPs as precursors, applying any synthesis
route previously mentioned, e.g., heating or applying plasma, to prepare the surface (which
may cause oxidation), coat it with polymers, and finally, carbonate it with an annealing
treatment [182–184].

2.4.3. Metallic Coatings


One of the disadvantages, if not the most important, of MNPs is the tendency to
change the crystal structure of the magnetite via oxidation [185]. The idea of linking a
nanoscale oxide phase with a metal surface to create a metal–oxide hybrid material with
unique features is appealing. Monodispersed iron oxide–metal nanostructures with binary
characteristics, such as core–shell, core–satellites, or dumbbell structures, achieve enhanced
stability and biocompatibility through the change in charge [186]. Noble metals in bulk
appear to have inert behavior, biocompatibility, and no significant interactions with other
elements, which could make them ideal in biomedical applications [187]. Due to their
interaction with external fields and the ease with which they are recovered, reused, or
altered with magnetic manipulation, MNPs, namely magnetite and maghemite, are particu-
larly appealing as components of multifunctional NMs, and the general biocompatibility
Processes 2022, 10, 2282 10 of 29

of noble metal nanoparticles makes these materials the ideal combination for biomedical
applications [186,188].
Conventionally, forming an amorphous or oxidate phase in nanostructured magnetic
materials has been a drawback of the synthesis process. Given its often negligible contribu-
tion to effective saturation magnetization, the amorphous fraction is frequently portrayed
as a “dead” magnetic component and a material waste [189]. The core–shell morphology
with an ordered core and a shell with structural/magnetic disorder often describes this
scenario as the simplest example of MNPs.

2.4.4. Polymer Coatings


Microemulsion polymerization and the sol–gel technique are frequently used in situ
synthesis methods that traditionally produce a core–shell or matrix-dispersed structure [49].
It is still difficult to regulate the colloidal stability and thickness of the created structures.
Post-synthesis methods such as the one-pot approach, self-assembly, or heterogeneous
polymerization are commonly chosen for the polymeric coating of MNPs to avoid excess
particle oxidation [190,191].
Utilizing polymer coatings has allowed for increased MNP stability and an expansion
of the application space by altering different functional groups [192] to provide properties
such as the connection with optoelectronic segments, imaging, diagnosis, and treatment,
enabling conjugated polymers with a delocalized electrical structure [193].
These are brief applications of covers and functionalization coatings that can be applied
to MNPs, known as multicoated or multicore. However, more properties can be attached
to these materials, combining coating types [194,195] to solve deficiencies of the iron NP
core, such as biocompatibility or recovery of the material. A typical method for creating
multicore nanoparticles is to trap many magnetic cores inside a polymeric matrix [195].
Three main approaches to the synthesis of polymer/SPION hybrid multicore particles have
been reported in the literature: (i) polymerization in the presence of the nanoparticles [196];
(ii) SPION formation from iron salts along an existing polymer particle; and (iii) in the
emulsification process (ESE) within a precipitating polymer [197].

2.5. Nanocomposites
Nanocomposites are materials with more than two phases, which have units with
at least one dimension within the nanoscale present in the matrix material to produce
integrated functional systems with additional properties [198]. Nanosized particles of
various magnetic materials have been inserted into extended matrix materials (i.e., organic
or inorganic polymers). Additionally, the future importance of magnetic nanocomposite
materials made from renewable resources will undoubtedly rise to safely recover the
material, making it a reusable source [199–201]. They blend the characteristics of the filler
with the matrix material to create innovative functional materials that are tailored to the
requirements of a specific application, such as when magnetite nanoparticles are enclosed
in a short-chain amphiphilic block copolymer combined with an organic pigment during
the self-assembly of higher-order polymer structures to create biosynthetic, bifunctional,
and magnetically active materials with poly(ethylene oxide)-block-poly(butadiene) (PEO-b-
PBD) [202].
In general, a wide variety of host materials, such as organic polymers, silica, or even
liquid media, are used in nanocomposites [203–205].

3. Biological Challenges
Nanotechnology, NMs, and nanostructures have been around humans since the early
days at the beginning of the industry and early metallurgy [206–208]. We can state that the
human–nanostructures interaction is not contemporary. Although the appropriate determi-
nation of “nano” was given centuries later by Richard Zsigmondy, the Nobel Laureate in
Chemistry (1925), he was the first to suggest the term “nanometer” [209] to the scientific
world. He was the first to measure the size of particles such as gold colloids [210–212],
Processes 2022, 10, 2282 11 of 29

and he developed the term nanometer to describe particle size using a microscope. How-
ever, only about 30 years ago, the significance of these nanostructured materials was
apparent to researchers of disciplines as diverse as medicine and photoelectronics due to
the different and novel characteristics conferred by their size. Researchers have discov-
ered that size affects materials’ physicochemical characteristics, such as their magnetic
properties [119,213,214].
MNPs can intervene in different steps and applications in both medical and industrial
sectors. Because these particles are composed of iron and oxygen, they have been targeted
as the standard material for medical applications because of their magnetic properties.
They can act as a carrier for drug and heat delivery [81,86,215,216], a potential treatment
when undergoing vibrating or carrier structure breaking processes [82,217–220], and even
a contrast and targeting agent [80,213,219–221], shortening periods between the diagnosis
and treatment, which is a crucial factor in medical processes and biocompatible needs. At
least, that was the original thought. Over the past two decades, research has shown mixed
results from the same tissue in different assays assessing the toxicity of MNPs (Table 2).
However, the concentration range is extensive. Most research focuses on the spherical
shape of the nanoparticle, giving controversial results depending on the study. Spherical
particles produce no effect when analyzed through ROS and TB in lung cells, but a comet
assay reveals oxidative DNA damage, as shown in Table 2.

Table 2. Different biological assays in MNPs *.

Tissue Concentration Morphology Size Coating Methodology Effect Ref.


Dextran,
Fibroblast
0.1–0.02 albumin
(hTERT Spherical 7.8–9.6 nm BrdU assay Cellular death [222–224]
mg/mL Lactoferrin,
human)
ceruloplasmin
Enhancement of
free radicals and
reduction in the
GSH
DNA oxidative
injuries (Comet)
Low—no toxicity
(TB, ROS)
Lung cells TB staining,
20–40 mg/kg Spherical 20–107.7 nm Bare Increased TP and [225,226]
(A549) ROS, Comet
LDH
Non-
biomechanical
damage
Cell Young’s
modulus
decreased
(25–28%)
Nontoxic below
75 µgmL−1
Nonalcoholic fatty
liver disease
20–40 mg/kg Bare
Liver rat cell 107.7 nm (NAFLD)
25, 50, 75, 150, Spherical Liposomes MTT, LDH [226–228]
(BAL3A rat) 20–30 nm inflammation
300 µg/g PEG
LDH leaking
Iron overload
affected by
NAFLD
Processes 2022, 10, 2282 12 of 29

Table 2. Cont.

Tissue Concentration Morphology Size Coating Methodology Effect Ref.


No significant
50, 100, effect
Mesenchymal Protamine
250 mg/mL 80–150 nm Increased
mother cell Spherical sulfate Comet [229–231]
25, 50, 100, 48 nm proliferation index
(MSC human) PDA
150 µg/mL and migration
ability
Cell viability
Phospholipid- reduced
Kidney cells based Particle charge
15 mg/kg Spheric-like 13–122 nm
(Cos-7 polymeric MTT, MTS (+)-induced high [232–234]
1–100 µg/mL Ferrofluid 9.7 nm
monkey) micelles cytotoxicity
DOX Oxidative stress,
reverted by tissue
Time-dependent
cell viability
Macrophage Ferrofluid Bare MTS, BrdU
2.73 mg/mL 320–490 nm (7 days, 20%) [235,236]
(human) Ellipsoidal SiO2 assay
Induced M1
activation
MTS increased
production
DNA
PGA
Nervous Spherical fragmentation,
SiO2
system cells 0–1000 µg/mL Hollow 5–100 nm MTS, LDH apoptotic [237–239]
Dextran
(human, PC12) spheres Conformational
Bare
changes in Tau
protein
Oxidative stress
Cellular intake
Cell viability
50 µg/mL 80 ± 3%
Endothelial 50–600 nm Bare
0–100 µg/mL Spherical PI staining, Promotes cell
cells (BAECs, 5–10 nm Dextran [240–242]
0, 300, Spherical–like Redox survival by
HUVECs) 10 nm PSC
600 µg/mL autophagy
Peroxidase-like
activity
High cell viability
>90%
Cancer 50, 100,
Spherical–like 6 nm DMSA MTT Trigger immune [240–243]
(multiple) 500 µg/mL
response
High ROS activity
Decreased cell
viability
Increased actin
and calponin
Vascular
50, 100, 200, Bare expression
system Spherical–like 150–160 nm Redox, MTT [242,244,245]
400 µg/mL Citric acid Concentration-
(A10 rat)
dependent
toxicity
Migration of EPC
reduced
* Table modified from [246]. BrdU—bromodeoxyuridine, TB—trypan blue, ROS—reactive oxygen species,
GSH—glutathione, TP—total protein, LDH—lactate dehydrogenase, MTT—3-[4,5-dimethylthiazole-2-yl]-2,5-
diphenyltetrazolium bromide, MTS—3-(4,5-dimethylthiazol-2-yl)-5-(3-carboxymethoxyphenyl)-2-(4-sulfophenyl)-
2H-tetrazolium, PI—propidium iodide, PEG—polyethylene glycol, PDA—polydopamine, DOX—doxorubicin,
PSC—polyglucose sorbitol carboxymethyl ether, DMSA—dimercaptosuccinic acid.

Although many studies have assessed the biological effects of MNPs, a consensus has
not been achieved on safety parameters.
Biological processes imply a series of pathways, reactions, and signaling at a molec-
ular level, all taking place around 1 to 100 nm, which is, coincidently, the action size
range of nanostructured materials. It is unsurprising then that, as such, nanotechnology
and nanostructure materials require special attention and safety parameters to function
ylene glycol, PDA—polydopamine, DOX—doxorubicin, PSC—polyglucose sorbitol carboxymethyl 
ether, DMSA—dimercaptosuccinic acid. 

Biological processes imply a series of pathways, reactions, and signaling at a molec‐
ular level, all taking place around 1 to 100 nm, which is, coincidently, the action size range 
Processes 2022, 10, 2282 13 of 29
of  nanostructured  materials.  It  is  unsurprising  then  that,  as  such,  nanotechnology  and 
nanostructure materials require special attention and safety parameters to function and 
not interfere with any other healthy cells and systems when applied to biological systems 
and not interfere with any other healthy cells and systems when applied to biological
(Figure 5). 
systems (Figure 5).

 
Figure 5. Size comparison of a red blood cell and a 70 nm nanoparticle.
Figure 5. Size comparison of a red blood cell and a 70 nm nanoparticle. 
Like any type of active drug substance and new material, nanostructured materials
must be subjected to different test methods to determine their effects and their impact on
health, regardless of whether they will be used in the medical sector or not [17,247–249].
 
Among the most used biological assays are the in vivo, in vitro, and the so-called ex
vivo tests. For the ex vivo tests, a target organ (heart, lungs, kidney, liver) is kept in a
controlled system outside the body to observe its interactions with an active ingredient
(such as nanoparticles or any vasoactive controls, acetylcholine, phenylephrine) in real
time [250–253]. Ex vivo tests have obvious advantages over other methods, including rapid
results, a relatively low cost, and reduced use of live animal exposure [254,255].
In the case of heavy metals, such as bulk iron, adverse effects at concentrations around
350 µgdL−1 are observed [256]. Additionally, the leaching of different ionic species can
occur in the ferrous state (Fe2+ ), which is insoluble, or in the ferric state (Fe3+ ), which can
directly interfere with agents related to oxidation [54,55,257]. Hemoglobin as molecular
oxygen (O2 ) and nitric oxide (NO2 , NO3 ), a representative mediator in cardiac muscle [258],
is a free radical synthesized enzymatically by endothelial NO synthase (eNOS), neuronal
synthases (nNOS), and inducible (iNOS) [258–261].
Detection of the non-favorable response of different substances in biological systems
can be classified into three main categories, depending on the biological section (cell,
tissue, organ, and system) that has been tested. In vitro is specifically for single cells and
tissue sections in controlled conditions in culture. In in vivo tests, the complex interactions
between systems (cardiovascular, respiratory, endocrine, digestive, renal, and nervous) are
tested. Moreover, a relatively new biological evaluation test has been raised, ex vivo, which
maintains isolated organs or segments in an external, controlled environment to detect
changes in physiology in real time (Figure 6).
In vivo studies can evaluate the primary stages of the adverse effects due to their
availability and relative ease of use compared to other studies. Ex vivo tests are a relatively
new approach to toxicology, bringing advantages over in vitro studies because of their
capacity to assess real-time effects on specific organs or tissue sections (heart, lungs, kidneys,
liver), isolating the tissue from the general bloodstream. Ex vivo analysis tends to isolate
specific organs from the animal and maintain them in a controlled environment. Complex
organs such as the heart require a more delicate approach in retrieval from the bloodstream,
necessitating a constant flow rate or pressure modes [262]; however, isolation from blocking
of the aorta and immediate recovery and attachment to the external parameters can preserve
the organ long enough to study it [250,253,263,264]. Lastly, in vivo assays are the last
Processes 2022, 10, 2282 14 of 29

Processes 2022, 10, x FOR PEER REVIEW  14  of  29 


  evaluation
before human trials and determine the complex response of different biological
system interactions and the impact of active substances.

 
6. Toxicological
Figure 6.
Figure  Toxicological assays and
assays  protocols
and  help
protocols  researchers
help  to better
researchers  understand
to  better  the impact
understand  of novel
the  impact  of 
novel pharmaceutical agents and materials: (a) in vitro, (b) ex vivo, and (c) in vivo. 
pharmaceutical agents and materials: (a) in vitro, (b) ex vivo, and (c) in vivo.

3.1. In Vitro Toxicology


3.1. In Vitro Toxicology 
Different in vitro tests are available to verify toxicity; the lactate dehydrogenases
Different in vitro tests are available to verify toxicity; the lactate dehydrogenases test 
test (LDH), sulforhodamine B (SRB) assay, protein assay, neutral red assay, and 3-(4,5-
(LDH), sulforhodamine B (SRB) assay, protein assay, neutral red assay, and 3‐(4,5‐dime‐
dimethylthiazol-2-yl)-2,5-diphenyltetrazolium
thylthiazol‐2‐yl)‐2,5‐diphenyltetrazolium  bromide  bromide (MTT)
(MTT)  assay
assay  are are a fewof ofthe 
a  few  theoften‐
often-
used assays [59].
used assays [59]. 
Tetrazolium salt, derived from MTT, is transformed into a purple, insoluble formazan
Tetrazolium salt, derived from MTT, is transformed into a purple, insoluble forma‐
complex by a mitochondrial dehydrogenase enzyme. These enzymes are only found in
zan complex by a mitochondrial dehydrogenase enzyme. These enzymes are only found 
active mitochondria; hence, the process only occurs in healthy cells [265]. The LDH leakage
in  active  mitochondria;  hence,  the  process  only  occurs  in  healthy  cells  [265].  The  LDH 
test, commonly used due to its dependability, speed, and ease of assessment, determines
leakage test, commonly used due to its dependability, speed, and ease of assessment, de‐
lactate dehydrogenase leakage to the extracellular medium in damaged cells [266]. Trypan
termines  lactate  dehydrogenase  leakage  to  the  extracellular  medium  in  damaged  cells 
blue is the test most often employed for viability studies. The test provides an uncompli-
[266]. Trypan blue is the test most often employed for viability studies. The test provides 
cated way to assess cellular viability by fixing the cells in a solution of trypan blue and
an uncomplicated way to assess cellular viability by fixing the cells in a solution of trypan 
paraformaldehyde after being deposited onto slides. Nonviable cells have a heavy blue
blue  and  paraformaldehyde  after  being  deposited  onto  slides.  Nonviable  cells  have  a 
stain, but viable cells do not [266]. However, this method does not distinguish between
heavy blue stain, but viable cells do not [266]. However, this method does not distinguish 
apoptotic cells.
between apoptotic cells. 
The safety assessment of MNPs on cell lines (in vitro) is uncomplicated, straightfor-
ward,The safety assessment of MNPs on cell lines (in vitro) is uncomplicated, straightfor‐
and affordable since the experiment can be reliably controlled. Toxicity in MNPs has
ward, and affordable since the experiment can be reliably controlled. Toxicity in MNPs 
been associated with dosage dependence, time, and surface modification [267].
has been associated with dosage dependence, time, and surface modification [267]. 
Since in vitro testing is the first approach to understanding the impact of material in a
livingSince in vitro testing is the first approach to understanding the impact of material in 
organism, it is not unusual that the information collected from in vitro and in vivo
a living organism, it is not unusual that the information collected from in vitro and in vivo 
tests happens to be controversial. This results from the complex bodily interactions and
tests happens to be controversial. This results from the complex bodily interactions and 
different systems coexisting in an in vivo test [268]. However, these tests unify the results
different systems coexisting in an in vivo test [268]. However, these tests unify the results 
of the full impact of the materials or pharmaceutical agents.
of the full impact of the materials or pharmaceutical agents. 
   
Processes 2022, 10, 2282 15 of 29

3.2. Ex Vivo Toxicity


Physiological models have been used for a long time as a valuable tool for the study
of endogenous molecules such as hormones, neurotransmitters, and other mediators under
normal or pathophysiological conditions, such as cardiovascular, respiratory, and digestive
disorders, to name a few [250,252].
Physiological ex vivo models of isolated tissues and organs allow the evaluation of the
particular functioning of a tissue or organ in NPs’ presence, for example, ducts related to
the cardiovascular system (aorta, arterioles, mesenteric, etc.), respiratory system (trachea,
bronchi, bronchioles), and digestive system (small or large intestine); or in the study of
organs such as the heart, kidney, lung, and liver; and the biochemical communications
between the organs and tissues [269]. By separating the organ from the general circulation,
identifying metabolites generated from the interaction with MNPs becomes easier [270].
This is a novel approach to assessing the toxicology of MNPs, but the road ahead is
still long and full of opportunity. These assays produce exciting results in cardiovascular
heart contractility and NO production.

3.3. In Vivo Toxicity


Although in vivo toxicity investigations take longer, cost more, entail ethical consider-
ations, and have more complicated techniques (such as toxicokinetic processes) [271], they
offer a clear benefit over in vitro experiments. Even though there is not enough in vivo
research on the toxic effects of MNPs, those studies that do exist provide valuable informa-
tion on possible toxicity, such as alternating current biosusceptometry (ACB) [272]. The
ACB system is a biomagnetic approach previously described and used in pharmacological
and gastrointestinal research in animals and people [273], as well as recently for detecting
MNPs [274].
For the best results in the therapeutic or diagnostic application, the MNPs are aggre-
gated in a specific tissue using a magnet, which might result in high concentrations in that
location [275]. This may result in large amounts of free Fe ions in the exposed tissue, result-
ing in cellular damage and/or significantly influencing subsequent cell generations [276].
When MNPs enter the body, they interact with the bodily fluids and then bind to
the proteins to form a protein corona, a nanoparticle–protein complex. The MNPs–corona
complex is essential for the particles to perform as intended because MNPs interact with
one another in vivo through the formation of this structure, which serves as the biological
identity of the MNPs [277]. As a result, understanding the surrounding protein composition,
particularly the individual proteins present in the MNPs–corona complex and their affinities,
is crucial for understanding how the MNPs will act in vivo.
These toxicological assays and approaches provide a different view of a material’s
effect on other parts of complex systems, increasing the understanding of its full impact on
nature and living organisms. All these efforts have expanded the body of knowledge on
nanotoxicology and help in regulating the novel properties that nanoscale materials bring
to different fields (Table 3).

Table 3. Biological assessment of treatments for MNPs.

Biological Studies Type of Assay Assays in MNPs Ref.


Suspension (HL60, K562) Monolayers (MCF-7,
In vitro U87MG) CCK8, MTT, TB, LDH, Comet [268,278]
Cultured
Perfusion pressure, protein
Ex vivo Langendorff isolated system, in silico studies expression, mediator count, [250,264,279]
liver, spleen, lungs, heart
In vivo Biodistribution, histological staining VIP, liver, spleen, lungs, heart [268,280]
CCK8—cell counting, VIP—vacuum filtration process.
Processes 2022, 10, 2282 16 of 29

4. Regulation and Control


Even though research and toxicity approaches have been made to identify the potential
toxicity of MNPs, there are no clear regulatory parameters to control these materials’
production, interaction, and disposal. Different organizations have made several proposals
since 2011 around the world (Table 4).

Table 4. Definitions of nanomaterials of different global organizations.

Organization Nanomaterial Definition * Status Last Meeting/Proposal Ref.


Taken from EC: “Any
intentionally produced material
that has one or more dimensions
NANoREG (European of the order of 100 nm or less or
Toxicological data 2014, updated by
Union, EU, European that is composed of discrete [281,282]
gathering NanoFATE in 2022
Commission, EC) functional parts, either internally
or at the surface, many of which
have one or more dimensions of
the order of 100 nm or less”.
“Any material with any external
International
dimension in the nanoscale or Terms and vocabulary
Organization for 2017 [283]
having an internal structure or for nano-objects
Standardization (ISO)
surface in the nanoscale”.
Taken from the NNI: “The
understanding and control of
FDA (United States of
matter at dimensions between Nonbinding
America, National
approximately 1 and recommendations for 2014 [284,285]
Nanotechnology
100 nanometers, where unique manufacturers
Initiative, NNI)
phenomena enable
novel applications”.
“A natural, incidental, or
manufactured material
containing particles, in an Guidance for terms,
unbound state or as an aggregate vocabulary, and sample
ECHA (European
or as an agglomerate and where, dispersion and Draft 2021 [286]
Union)
for 50% or more of the particles aggregation of
in the number size distribution, nanoforms
one or more external dimensions
is in size range 1 nm–100 nm”.
“Any manufactured substance or
product, as well as any
component material, ingredient,
device, or structure, if it has at
least one external dimension that
Guidance framework
is at or within the nanoscale, or
for adapting
CEPA (Canada) if it has internal or surface Draft 2022 [287]
nanomaterials to
structure that is at the nanoscale,
existing practices
or if it has all dimensions that
are smaller or larger than the
nanoscale and exhibits at least
one nanoscale property
or phenomenon”.
* The definition was extracted literally from each organization manuscript cited in the reference column.

The idea of safe-by-design is becoming an increasingly significant approach to nanosafety.


Nanotechnology-based products must be safe to use throughout their lifespan, from manufac-
ture to disposal, recycling, and reuse. A well-established general idea in industrial innovation,
safe-by-design was first developed for nanomaterials in the European Union’s (EU) flagship
project NANoREG [281].
Processes 2022, 10, 2282 17 of 29

The legislation on Registration, Evaluation, Authorization, and Restriction of Chemi-


cals, or REACH, in the European Union is responsible for regulating chemical agents [288].
REACH defines any NMs as “nanoforms,” creating a new guideline with which to contem-
plate these new materials and their properties. It has a section establishing that a material
composed of less than 50% nanostructures is classified as a bulk material without consider-
ing its dissemination to the environment or biological organisms. However, these guidelines
were defined in 2013, and just recently, the term nanoform was implemented [286].
Another administration agency, the U.S. Food and Drug Administration (FDA), which
is in charge of the regulation of product safety, has provided a guideline to specify if a
manufactured good possesses nanotechnology or nanoscale properties with two criteria:
(i) whether a substance or finished product is designed to have at least one exterior dimen-
sion, an interior structure, or a surface structure in the nanoscale range (approximately
1 nm to 100 nm) and (ii) whether a substance or finished product is developed to show
characteristics or phenomena, such as physical or chemical characteristics or biological con-
sequences, even if these dimensions fall beyond the nanoscale range, up to one micrometer
(1000 nm) [285]. The FDA guideline, however, was released in 2014 and concluded that
nanomaterials are not subject to premarket review but urged the industry to consult with
the agency early in the product development process, leaving nanomaterials stranded with
no regulations.
Canada’s regulations on environmental and health hazards are controlled by the Cana-
dian Environmental Protection Act (CEPA), which put out regulatory parameters in 2013 on
how to modify current risk assessment procedures considering the unique features shown
by chemicals at the nanoscale in line with the Organization for Economic Co-operation
and Development (OECD) proposal [287]. In the latest draft of June 2022, the CEPA pro-
posed mediating NM risk in three appraisals: (i) substance identity, taking into account
composition, shape, agglomeration, surface chemistry, dissolution in biological media,
and any other specification that the material may possess according to the manufacturer;
(ii) exposure assessment, be it direct contact or indirect contact (water, air, soil), and release
potential or fragmentation; (iii) hazard assessment, considering toxicological endpoints
such as acute toxicity, repeated-dose toxicity, genetic toxicity, reproductive or developmen-
tal toxicity, carcinogenicity, toxicokinetics, etc. The final step is the risk characterization, as
determined by the points established in part 5, Section 64 of CEPA: (a) the material has or
may have an immediate or long-term harmful effect on the environment or its biological
diversity; (b) the material constitutes or may constitute a danger to the environment on
which life depends; and (c) the material constitutes or may constitute a danger in Canada
to human life or health.
It is well-known that MNPs have different properties than their counterparts on a
macroscopic scale. Studies have been conducted to determine their toxicity in various body
organs with multiple materials [289–291]. However, reported viability results on the same
material in different tissues within the body have presented contradictory results depending
on the type of tissue to which these materials were administrated [1,51,59,219,248,292–294],
so their regulation must be specialized. Furthermore, it is marked correctly to prevent a
significant effect on society and the environment [295].

5. Conclusions and Perspective


MNPs exhibit interesting novel properties due to their superparamagnetism, reduced
size (10 nm), and versatility in forming functionalized composites that increase their appli-
cations as theranostics. However, these same problem-solving qualities can be responsible
for causing possible adverse biological impacts in the future if they result in excessive
and unregulated use. Different global initiatives have been launched to address this new
branch called nanotoxicology and nanoethics. This must be considered when applying
any NMs in any sector, not just MNPs, whether it is industry or medicine, to ensure the
correct handling of the material and prevent other environmental and health problems due
to the improper use of new versatile materials. However, the vast properties of NPs, even
Processes 2022, 10, 2282 18 of 29

with the same composition, be it size, morphology, surface charge, magnetism, or general
reactivity, present a problematic task ahead in mitigating NPs’ effects. Slowly but steadily,
though, the regulations for MNPs are taking shape. Still, the interaction and pathways of
NPs are yet to be uncovered. There is no comprehensive image of the toxicological impact
of magnetic nanoparticles on biological cells. Even though the literature is full of extensive
analyses of different biological systems discussing the toxicology aspects of MNPs used in
biomedical applications, they give mixed and inconclusive results depending on the assay
implemented to determine their biological effects with an exhaustive list of parameters.
Narrowing this down to a consensus list would help to generate a consolidative framework
for these biological behaviors to allow manageable future applications. Finally, using only
one nanomaterial parameter in the same tissue to make a consensus on the adverse effects
and impact of MNPs leads to unreliable results.

Author Contributions: Conceptualization, J.R.V.-O., C.G. and K.E.; investigation, J.R.V.-O.; resources,
C.G. and K.E.; writing—original draft preparation, J.R.V.-O.; writing—review and editing, C.G. and
K.E.; visualization, C.G. and K.E.; supervision, K.E.; project administration, C.G. and K.E.; funding
acquisition, K.E. All authors have read and agreed to the published version of the manuscript.
Funding: This research was funded by Universidad Autónoma de Querétaro (FIN202116; FIN202106).
Data Availability Statement: Not applicable.
Acknowledgments: JRVO thanks CONACyT for the scholarship granted. KE thanks the engineering
faculty at UAQ for the financial support given through the Attention to National Problems fund
FI-UAQ FIN202116 and the Universidad Autónoma de Querétaro through the fund FONDEC-UAQ
2021FIN202106. All the images were created with BioRender.com (PG24DGDY7O; RY24DGDYA6;
EC24DGDYC5; PU24DGDYEK).
Conflicts of Interest: The authors declare no conflict of interest. The funders had no role in the design
of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or
in the decision to publish the results.

References
1. Khan, I.; Saeed, K.; Khan, I. Nanoparticles: Properties, applications and toxicities. Arab. J. Chem. 2019, 12, 908–931. [CrossRef]
2. Missaoui, W.N.; Arnold, R.D.; Cummings, B.S. Toxicological status of nanoparticles: What we know and what we don’t know.
Chem. Biol. Interact. 2018, 295, 1–12. [CrossRef] [PubMed]
3. Scarpelli, F.; Mastropietro, T.F.; Poerio, T.; Godbert, N. Mesoporous TiO2 Thin Films: State of the Art. Titan. Dioxide-Mater. A
Sustain. Environ. 2018, 508, 135–142.
4. Madkour, L.H. Environmental Impact of Nanotechnology and Novel Applications of Nano Materials and Nano Devices; Springer: Cham,
Germany, 2019; Volume 116.
5. Prabha, S.; Arya, G.; Chandra, R.; Ahmed, B.; Nimesh, S. Effect of size on biological properties of nanoparticles employed in gene
delivery. Artif. Cells Nanomed. Biotechnol. 2016, 44, 83–91. [CrossRef]
6. Gong, S.; Cheng, W. One-Dimensional Nanomaterials for Soft Electronics. Adv. Electron. Mater. 2017, 3, 1600314. [CrossRef]
7. Navalón, S.; García, H. Nanoparticles for catalysis. Nanomaterials 2016, 6, 123. [CrossRef]
8. Turci, F.; Pavan, C.; Leinardi, R.; Tomatis, M.; Pastero, L.; Garry, D.; Anguissola, S.; Lison, D.; Fubini, B. Revisiting the paradigm of
silica pathogenicity with synthetic quartz crystals: The role of crystallinity and surface disorder. Part. Fibre Toxicol. 2016, 13, 1–12.
[CrossRef]
9. Selim, A.A.; Al-Sunaidi, A.; Tabet, N. Effect of the surface texture and crystallinity of ZnO nanoparticles on their toxicity. Mater.
Sci. Eng. C 2012, 32, 2356–2360. [CrossRef]
10. Usov, N.A.; Rytov, R.A.; Bautin, V.A. Properties of assembly of superparamagnetic nanoparticles in viscous liquid. Sci. Rep. 2021,
11, 1–11.
11. Hu, M.; Butt, H.-J.; Landfester, K.; Bannwarth, M.B.; Wooh, S.; Thérien-Aubin, H. Shaping the Assembly of Superparamagnetic
Nanoparticles. ACS Nano 2019, 13, 3015–3022. [CrossRef]
12. Yin, X.; Russek, S.E.; Zabow, G.; Sun, F.; Mohapatra, J.; Keenan, K.E.; Boss, M.A.; Zeng, H.; Liu, J.P.; Viert, A.; et al. Large T 1
contrast enhancement using superparamagnetic nanoparticles in ultra-low field MRI. Sci. Rep. 2018, 8, 1–10. [CrossRef] [PubMed]
13. Szpak, A.; Fiejdasz, S.; Prendota, W.; Straczek,
˛ T.; Kapusta, C.; Szmyd, J.; Nowakowska, M.; Zapotoczny, S. T1–T2 Dual-modal
MRI contrast agents based on superparamagnetic iron oxide nanoparticles with surface attached gadolinium complexes. J.
Nanoparticle Res. 2014, 16, 1–11. [CrossRef] [PubMed]
14. Gao, Y.; Shi, X.; Shen, M. Intelligent Design of Ultrasmall Iron Oxide Nanoparticle-Based Theranostics. ACS Appl. Mater. Interfaces
2021, 13, 45119–45129. [CrossRef] [PubMed]
Processes 2022, 10, 2282 19 of 29

15. Chaves, N.L.; Estrela-Lopis, I.; Böttner, J.; Lopes, C.A.P.; Guido, B.C.; de Souza, A.R.; Báo, S.N. Exploring cellular uptake of iron
oxide nanoparticles associated with rhodium citrate in breast cancer cells. Int. J. Nanomed. 2017, 12, 5511–5523. [CrossRef]
16. Shrestha, S.; Wang, B.; Dutta, P. Nanoparticle processing: Understanding and controlling aggregation. Adv. Colloid Interface Sci.
2020, 279, 102162. [CrossRef]
17. Kendall, M.; Ding, P.; Kendall, K. Particle and nanoparticle interactions with fibrinogen: The importance of aggregation in
nanotoxicology. Nanotoxicology 2010, 5, 55–65. [CrossRef]
18. Babakhani, P. The impact of nanoparticle aggregation on their size exclusion during transport in porous media: One- and
three-dimensional modelling investigations. Sci. Rep. 2019, 9, 1–12. [CrossRef]
19. Ashraf, M.A.; Peng, W.; Zare, Y.; Rhee, K.Y. Effects of Size and Aggregation/Agglomeration of Nanoparticles on the Interfa-
cial/Interphase Properties and Tensile Strength of Polymer Nanocomposites. Nanoscale Res. Lett. 2018, 13, 1–7. [CrossRef]
20. Bahadar, H.; Maqbool, F.; Niaz, K.; Abdollahi, M. Toxicity of Nanoparticles and an Overview of Current Experimental Models.
Iran. Biomed. J. 2016, 20, 1–11.
21. Magdolenova, Z.; Collins, A.; Kumar, A.; Dhawan, A.; Stone, V.; Dusinska, M. Mechanisms of genotoxicity. A review of in vitro
and in vivo studies with engineered nanoparticles. Nanotoxicology 2014, 8, 233–278. [CrossRef]
22. Sun, D.; Gong, L.; Xie, J.; Gu, X.; Li, Y.; Cao, Q.; Li, Q.; Luodan, A.; Gu, Z.; Xu, H. Toxicity of silicon dioxide nanoparticles with
varying sizes on the cornea and protein corona as a strategy for therapy. Sci. Bull. 2018, 63, 907–916. [CrossRef]
23. Pope, C.A.; Cohen, A.J.; Burnett, R.T. Cardiovascular disease and fine particulate matter lessons and limitations of an integrated
exposure-response approach. Circ. Res. 2018, 122, 1645–1647. [CrossRef] [PubMed]
24. United States Environmental Protection Agency. Particle Pollution and Cardiovascular Effects. 2021. Available online:
https://www.epa.gov/pmcourse/particle-pollution-and-cardiovascular-effects (accessed on 24 February 2022).
25. Schulz, H.; Harder, V.; Ibald-Mulli, A.; Khandoga, A.; Koenig, W.; Krombach, F.; Radykewicz, R.; Stampfl, A.; Thorand, B.;
Peters, A. Cardiovascular Effects of Fine and Ultrafine Particles. J. Aerosol Med. 2005, 18, 1–22. [CrossRef] [PubMed]
26. Hoshyar, N.; Gray, S.; Han, H.; Bao, G. The effect of nanoparticle size on in vivo pharmacokinetics and cellular interaction.
Nanomedicine 2016, 11, 673–692. [CrossRef]
27. Savliwala, S.; Chiu-Lam, A.; Unni, M.; Rivera-Rodriguez, A.; Fuller, E.; Sen, K.; Threadcraft, M.; Rinaldi, C. Magnetic Nanoparticles;
Elsevier Inc.: Amsterdam, The Netherlands, 2019.
28. FDA. Information on Gadolinium-Based Contrast Agents Regulatory History and Labeling from Drugs @ FDA; FDA:
Silver Spring, MD, USA, 2017.
29. Xiao, Y.-D.; Paudel, R.; Liu, J.; Ma, C.; Zhang, Z.-S.; Zhou, S.-K. MRI contrast agents: Classification and application (Review). Int.
J. Mol. Med. 2016, 38, 1319–1326. [CrossRef]
30. Zhen, Z.; Xie, J. Development of Manganese-Based Nanoparticles as Contrast Probes for Magnetic Resonance Imaging. Theranostics
2012, 2, 45–54. [CrossRef]
31. De León-Rodríguez, L.M.; Martins, A.F.; Pinho, M.C.; Rofsky, N.M.; Sherry, A.D. Basic MR relaxation mechanisms and contrast
agent design. J. Magn. Reson. Imaging 2015, 42, 545–565. [CrossRef]
32. Neeley, C.; Moritz, M.; Brown, J.J.; Zhou, Y. Acute side effects of three commonly used gadolinium contrast agents in the paediatric
population. Br. J. Radiol. 2016, 89, 20160027. [CrossRef]
33. Watson, A.D. The use of gadolinium and dysprosium chelate complexes as contrast agents for magnetic resonance imaging This
substituent group is believed to provide the required. J. Alloy. Compd. 1994, 207, 14–19. [CrossRef]
34. Norek, M.; Peters, J.A. MRI contrast agents based on dysprosium or holmium. Prog. Nucl. Magn. Reson. Spectrosc. 2011, 59, 64–82.
[CrossRef]
35. Urian, Y.; Atoche-Medrano, J.; Quispe, L.T.; Félix, L.L.; Coaquira, J. Study of the surface properties and particle-particle interactions
in oleic acid-coated Fe3O4 nanoparticles. J. Magn. Magn. Mater. 2021, 525, 167686. [CrossRef]
36. Kaur, I.P.; Kakkar, V.; Deol, P.K.; Yadav, M.; Singh, M.; Sharma, I. Issues and concerns in nanotech product development and its
commercialization. J. Control Release 2014, 193, 51–62. [CrossRef] [PubMed]
37. Deng, L.; Liu, Z.; Li, L. Hybrid nanocomposites for imaging-guided synergistic theranostics. In Nanomaterials for Drug Delivery
and Therapy; Elsevier Inc.: Amsterdam, The Netherlands, 2019; pp. 117–147.
38. Oliveira, E.; Rocha, M.; Froner, A.P.; Basso, N.; Zanini, M.; Papaléo, R. Synthesis and nuclear magnetic relaxation properties of
composite iron oxide nanoparticles. Quim. Nova 2018, 42, 57–64. [CrossRef]
39. Williams, H.M. The application of magnetic nanoparticles in the treatment and monitoring of cancer and infectious diseases.
Biosci. Horizons 2017, 10, 1–10. [CrossRef]
40. Lamon, L.; Asturiol, D.; Richarz, A.; Joossens, E.; Graepel, R.; Aschberger, K.; Worth, A. Grouping of nanomaterials to read-across
hazard endpoints: From data collection to assessment of the grouping hypothesis by application of chemoinformatic techniques.
Part. Fibre Toxicol. 2018, 15, 1–17. [CrossRef]
41. Hensley, D.; Tay, Z.W.; Dhavalikar, R.; Zheng, B.; Goodwill, P.; Rinaldi, C.; Conolly, S. Combining magnetic particle imaging and
magnetic fluid hyperthermia in a theranostic platform. Phys. Med. Biol. 2017, 62, 3483–3500. [CrossRef]
42. Hapuarachchige, S.; Artemov, D. Theranostic Pretargeting Drug Delivery and Imaging Platforms in Cancer Precision Medicine.
Front. Oncol. 2020, 10, 1131. [CrossRef]
43. Thorat, N.D.; Lemine, O.M.; Bohara, R.A.; Omri, K.; El Mir, L.; Tofail, S.A.M. Superparamagnetic iron oxide nanocargoes for
combined cancer thermotherapy and MRI applications. Phys. Chem. Chem. Phys. 2016, 18, 21331–21339. [CrossRef]
Processes 2022, 10, 2282 20 of 29

44. Siddhardha, B.; Parasuraman, P. Theranostics application of nanomedicine in cancer detection and treatment. In Nanomaterials for
Drug Delivery and Therapy; Elsevier Inc.: Amsterdam, The Netherlands, 2019; pp. 59–89.
45. Kosuda, K.M.; Bingham, J.M.; Wustholz, K.L.; Van Duyne, R.P. Nanostructures and Surface-Enhanced Raman Spectroscopy. In
Handbook of Nanoscale Optics and Electronicsvol; Elsevier Ltd.: Amsterdam, The Netherlands, 2010; Volume 1–5.
46. Morcos, B.; Lecante, P.; Morel, R.; Haumesser, P.H.; Santini, C.C. Magnetic, structural and chemical properties of cobalt
nanoparticles synthesized in ionic liquids Bishoy. Langmuir 2018, 34, 7086–7095. [CrossRef]
47. Ahghari, M.R.; Soltaninejad, V.; Maleki, A. Synthesis of nickel nanoparticles by a green and convenient method as a magnetic
mirror with antibacterial activities. Sci. Rep. 2020, 10, 1–10. [CrossRef]
48. Malhotra, N.; Lee, J.-S.; Liman, R.A.D.; Ruallo, J.M.S.; Villaflores, O.B.; Ger, T.-R.; Hsiao, C.-D. Potential Toxicity of Iron Oxide
Magnetic Nanoparticles: A Review. Molecules 2020, 25, 3159. [CrossRef] [PubMed]
49. Patsula, V.; Tulinska, J.; Trachtová, Š.; Kuricova, M.; Liskova, A.; Španová, A.; Ciampor, F.; Vavra, I.; Rittich, B.; Ursinyova, M.
Toxicity evaluation of monodisperse PEGylated magnetic nanoparticles for nanomedicine. Nanotoxicology 2019, 13, 510–526.
[CrossRef] [PubMed]
50. Genevière, A.-M.; Derelle, E.; Escande, M.-L.; Grimsley, N.; Klopp, C.; Ménager, C.; Michel, A.; Moreau, H. Responses to iron
oxide and zinc oxide nanoparticles in echinoderm embryos and microalgae: Uptake, growth, morphology, and transcriptomic
analysis. Nanotoxicology 2020, 14, 1342–1361. [CrossRef]
51. Guggenheim, E.J.; Rappoport, J.Z.; Lynch, I. Mechanisms for cellular uptake of nanosized clinical MRI contrast agents. Nanotoxi-
cology 2020, 14, 504–532. [CrossRef] [PubMed]
52. Feng, Q.; Liu, Y.; Huang, J.; Chen, K.; Huang, J.; Xiao, K. Uptake, distribution, clearance, and toxicity of iron oxide nanoparticles
with different sizes and coatings. Sci. Rep. 2018, 8, 1–13. [CrossRef]
53. Zhao, J.; Brugger, J.; Pring, A. Mechanism and kinetics of hydrothermal replacement of magnetite by hematite. Geosci. Front. 2019,
10, 29–41. [CrossRef]
54. Qiu, T.-S.; Fang, X.-H.; Wu, H.-Q.; Zeng, Q.-H.; Zhu, D.-M. Leaching behaviors of iron and aluminum elements of ion-absorbed-
rare-earth ore with a new impurity depressant. Trans. Nonferrous Met. Soc. China 2014, 24, 2986–2990. [CrossRef]
55. Strasser, H.; Brunner, H.; Schinner, F. Leaching of iron and toxic heavy metals from anaerobically-digested sewage sludge. J. Ind.
Microbiol. Biotechnol. 1995, 14, 281–287. [CrossRef]
56. Polasky, C.; Studt, T.; Steuer, A.-K.; Loyal, K.; Lüdtke-Buzug, K.; Bruchhage, K.-L.; Pries, R. Impact of Superparamagnetic Iron
Oxide Nanoparticles on THP-1 Monocytes and Monocyte-Derived Macrophages. Front. Mol. Biosci. 2022, 9. [CrossRef]
57. Singh, N.; Jenkins, G.J.; Asadi, R.; Doak, S.H. Potential toxicity of superparamagnetic iron oxide nanoparticles (SPION). Nano Rev.
2010, 1, 5358. [CrossRef]
58. Du, S.; Li, J.; Du, C.; Huang, Z.; Chen, G.; Yan, W. Overendocytosis of superparamagnetic iron oxide particles increases apoptosis
and triggers autophagic cell death in human osteosarcoma cell under a spinning magnetic field. Oncotarget 2016, 8, 9410–9424.
[CrossRef] [PubMed]
59. Patil, R.M.; Thorat, N.D.; Shete, P.B.; Bedge, P.A.; Gavde, S.; Joshi, M.G.; Tofail, S.A.; Bohara, R.A. Comprehensive cytotoxicity
studies of superparamagnetic iron oxide nanoparticles. Biochem. Biophys. Rep. 2018, 13, 63–72. [CrossRef] [PubMed]
60. Jiang, P.; Gan, M.; Yen, S.-H.; Dickson, D.W. Nanoparticles With Affinity for α-Synuclein Sequester α-Synuclein to Form Toxic
Aggregates in Neurons With Endolysosomal Impairment. Front. Mol. Neurosci. 2021, 14, 1–14. [CrossRef]
61. Shukla, R.K.; Badiye, A.; Vajpayee, K.; Kapoor, N. Genotoxic Potential of Nanoparticles: Structural and Functional Modifications
in DNA. Front. Genet. 2021, 12, 1–16. [CrossRef] [PubMed]
62. Russell, E.; Dunne, V.; Russell, B.; Mohamud, H.; Ghita, M.; McMahon, S.J.; Butterworth, K.T.; Schettino, G.; McGrry, C.K.;
Prise, K.M. Impact of superparamagnetic iron oxide nanoparticles on in vitro and in vivo radiosensitisation of cancer cells. Radiat.
Oncol. 2021, 16, 1–16. [CrossRef] [PubMed]
63. Cellai, F.; Munnia, A.; Viti, J.; Doumett, S.; Ravagli, C.; Ceni, E.; Mello, T.; Polvani, S.; Giese, R.W.; Baldi, G.; et al. Magnetic
Hyperthermia and Oxidative Damage to DNA of Human Hepatocarcinoma Cells. Int. J. Mol. Sci. 2017, 18, 939. [CrossRef]
[PubMed]
64. Döpke, C.; Grothe, T.; Steblinski, P.; Klöcker, M.; Sabantina, L.; Kosmalska, D.; Blachowicz, T.; Ehrmann, A. Magnetic Nanofiber
Mats for Data Storage and Transfer. Nanomaterials 2019, 9, 92. [CrossRef]
65. Grothe, T.; Sabantina, L.; Klöcker, M.; Junger, I.J.; Döpke, C.; Ehrmann, A. Wet Relaxation of Electrospun Nanofiber Mats.
Technologies 2019, 7, 23. [CrossRef]
66. Papavasileiou, A.; Panagiotopoulos, I.; Prodromidis, M.I. All-screen-printed graphite sensors integrating permanent bonded
magnets. Fabrication, characterization and analytical utility. Electrochimica Acta 2020, 360, 136981. [CrossRef]
67. Verma, S.; Kumar, V.; Gupta, K.D. Performance analysis of flexible multirecess hydrostatic journal bearing operating with
micropolar lubricant. Lubr. Sci. 2012, 24, 273–292. [CrossRef]
68. Shahidi, S. Magnetic nanoparticles application in the textile industry—A review. J. Ind. Text. 2019, 50, 970–989. [CrossRef]
69. Bustamante-Torres, M.; Romero-Fierro, D.; Arcentales-Vera, B.; Pardo, S.; Bucio, E. Interaction between Filler and Polymeric
Matrix in Nanocomposites: Magnetic Approach and Applications. Polymers 2021, 13, 2998. [CrossRef] [PubMed]
70. Coutinho, M.; Miranda, J.A. Peak instability in an elastic interface ferrofluid. Phys. Fluids 2020, 32, 5. [CrossRef]
Processes 2022, 10, 2282 21 of 29

71. Peyghami, A.; Moharrami, A.; Rashtbari, Y.; Afshin, S.; Vosuoghi, M.; Dargahi, A. Evaluation of the efficiency of magnetized
clinoptilolite zeolite with Fe3 O4 nanoparticles on the removal of basic violet 16 (BV16) dye from aqueous solutions. J. Dispers. Sci.
Technol. 2021, 1–10. [CrossRef]
72. Dargahi, A.; Hasani, K.; Mokhtari, S.A.; Vosoughi, M.; Moradi, M.; Vaziri, Y. Highly effective degradation of 2,4-
Dichlorophenoxyacetic acid herbicide in a three-dimensional sono-electro-Fenton (3D/SEF) system using powder activated
carbon (PAC)/Fe3O4 as magnetic particle electrode. J. Environ. Chem. Eng. 2021, 9, 105889. [CrossRef]
73. Seidmohammadi, A.; Vaziri, Y.; Dargahi, A.; Nasab, H.Z. Improved degradation of metronidazole in a heterogeneous photo-
Fenton oxidation system with PAC/Fe3O4 magnetic catalyst: Biodegradability, catalyst specifications, process optimization, and
degradation pathway. Biomass Convers. Biorefinery 2021, 1–17. [CrossRef]
74. Seabra, A.B.; Pelegrino, M.T.; Haddad, P.S. Antimicrobial Applications of Superparamagnetic Iron Oxide Nanoparticles: Perspectives and
Challenges; Elsevier Inc.: Amsterdam, The Netherlands, 2017.
75. Blachowicz, T.; Ehrmann, A. Magnetization reversal in bent nanofibers of different cross sections. J. Appl. Phys. 2018, 124, 152112.
[CrossRef]
76. Chaparro, C.M.; Suchdev, P.S. Anemia epidemiology, pathophysiology, and etiology in low- and middle-income countries. Ann.
N. Y. Acad. Sci. 2019, 1450, 15–31. [CrossRef]
77. Elshemy, M.A. Iron Oxide Nanoparticles Versus Ferrous Sulfate In Treatment of Iron Deficiency Anemia In Rats. Egypt. J. Vet. Sci.
2018, 49, 103–109. [CrossRef]
78. Wang, A.; Bagalkot, V.; Vasilliou, C.C.; Gu, F.; Alexis, F.; Zhang, L.; Shaikh, M.; Yuet, K.; Cima, M.J.; Langer, R.; et al. Superparam-
agnetic Iron Oxide Nanoparticle-Aptamer Bioconjugates for Combined Prostate Cancer Imaging and Therapy. ChemMedChem
2008, 3, 1311–1315. [CrossRef]
79. Harrison, R.J.; Dunin-Borkowski, R.E.; Putnis, A. Direct imaging of nanoscale magnetic interactions in minerals. Proc. Natl. Acad.
Sci. USA 2002, 99, 16556–16561. [CrossRef] [PubMed]
80. Wáng, Y.X.J.; Idée, J.M. A comprehensive literatures update of clinical researches of superparamagnetic resonance iron oxide
nanoparticles for magnetic resonance imaging. Quant. Imaging Med. Surg. 2017, 7, 88–122. [CrossRef] [PubMed]
81. Thorat, N.D.; Bohara, R.A.; Malgras, V.; Tofail, S.A.M.; Ahamad, T.; Alshehri, S.M.; Wu, K.C.-W.; Yamauchi, Y. Multimodal
Superparamagnetic Nanoparticles with Unusually Enhanced Specific Absorption Rate for Synergetic Cancer Therapeutics and
Magnetic Resonance Imaging. ACS Appl. Mater. Interfaces 2016, 8, 14656–14664. [CrossRef]
82. Jouyandeh, M.; Paran, S.M.R.; Shabanian, M.; Ghiyasi, S.; Vahabi, H.; Badawi, M.; Formela, K.; Puglia, D.; Saeb, M.R. Cur-
ing behavior of epoxy/Fe3O4nanocomposites: A comparison between the effects of bare Fe3O4, Fe3O4/SiO2/chitosan and
Fe3O4/SiO2/chitosan/imide/phenylalanine-modified nanofillers. Prog. Org. Coat. 2018, 123, 10–19. [CrossRef]
83. Darwish, M.; Kim, H.; Bui, M.; Le, T.-A.; Lee, H.; Ryu, C.; Lee, J.; Yoon, J. The Heating Efficiency and Imaging Performance of
Magnesium Iron Oxide@tetramethyl Ammonium Hydroxide Nanoparticles for Biomedical Applications. Nanomaterials 2021,
11, 1096. [CrossRef] [PubMed]
84. Stueber, D.; Villanova, J.; Aponte, I.; Xiao, Z.; Colvin, V. Magnetic Nanoparticles in Biology and Medicine: Past, Present, and
Future Trends. Pharmaceutics 2021, 13, 943. [CrossRef]
85. Xu, L.; Zhong, S.; Shi, C.; Sun, Y.; Zhao, S.; Gao, Y.; Cui, X. Sonochemical fabrication of reduction-responsive magnetic starch-based
microcapsules. Ultrason. Sonochem. 2018, 49, 169–174. [CrossRef]
86. Kunrath, M.F.; Campos, M.M. Metallic-nanoparticle release systems for biomedical implant surfaces: Effectiveness and safety.
Nanotoxicology 2021, 15, 721–739. [CrossRef]
87. Hu, T.; Mei, X.; Wang, Y.; Weng, X.; Liang, R.; Wei, M. Two-dimensional nanomaterials: Fascinating materials in biomedical field.
Sci. Bull. 2019, 64, 1707–1727. [CrossRef]
88. Gualdani, R.; Guerrini, A.; Fantechi, E.; Tadini-Buoninsegni, F.; Moncelli, M.R.; Sangregorio, C. Superparamagnetic iron oxide
nanoparticles (SPIONs) modulate hERG ion channel activity. Nanotoxicology 2019, 13, 1197–1209. [CrossRef]
89. Tian, F.; Chen, G.; Yi, P.; Zhang, J.; Li, A.; Zhang, J.; Zheng, L.; Deng, Z.; Shi, Q.; Peng, R.; et al. Fates of Fe3 O4 and
Fe3 O4 @SiO2 nanoparticles in human mesenchymal stem cells assessed by synchrotron radiation-based techniques. Bioma-
terials 2014, 35, 6412–6421. [CrossRef] [PubMed]
90. Carmona-Carmona, A.J.; Palomino-Ovando, M.A.; Hernández-Cristobal, O.; Sánchez-Mora, E.; Toledo-Solano, M. Synthesis and
characterization of magnetic opal/Fe3O4 colloidal crystal. J. Cryst. Growth 2017, 462, 6–11. [CrossRef]
91. Awada, H.; Al Samad, A.; Laurencin, D.; Gilbert, R.; Dumail, X.; El Jundi, A.; Bethry, A.; Pomrenke, R.; Johnson, C.; Lemaire, L.; et al.
Controlled Anchoring of Iron Oxide Nanoparticles on Polymeric Nanofibers: Easy Access to Core@Shell Organic–Inorganic Nanocom-
posites for Magneto-Scaffolds. ACS Appl. Mater. Interfaces 2019, 11, 9519–9529. [CrossRef]
92. Yazid, N.A.; Joon, Y.C. Co-precipitation synthesis of magnetic nanoparticles for efficient removal of heavy metal from synthetic
wastewater Co-precipitation Synthesis of Magnetic Nanoparticles for Efficient Removal of Heavy Metal from Synthetic Wastewater.
In AIP Conference Proceedings; AIP Publishing LLC: Melville, NY, USA, 2019; Volume 2124, p. 020019.
93. Daoush, W.M. Co-Precipitation and Magnetic Properties of Magnetite Nanoparticles for Potential Biomedical Applications. J.
Nanomed. Res. 2017, 5, 1–6. [CrossRef]
94. Mohammadi, H.; Nekobahr, E.; Akhtari, J.; Saeedi, M.; Akbari, J.; Fathi, F. Synthesis and characterization of magnetite nanoparti-
cles by co-precipitation method coated with biocompatible compounds and evaluation of in-vitro cytotoxicity. Toxicol. Rep. 2021,
8, 331–336. [CrossRef] [PubMed]
Processes 2022, 10, 2282 22 of 29

95. Cotin, G.; Kiefer, C.; Perton, F.; Ihiawakrim, D.; Blanco-Andujar, C.; Moldovan, S.; Lefevre, C.; Ersen, O.; Pichon, B.; Mertz, D.; et al.
Unravelling the Thermal Decomposition Parameters for The Synthesis of Anisotropic Iron Oxide Nanoparticles. Nanomaterials
2018, 8, 881. [CrossRef]
96. Unni, M.; Uhl, A.M.; Savliwala, S.; Savitzky, B.H.; Dhavalikar, R.; Garraud, N.; Arnold, D.P.; Kourkoutis, L.F.; Andrew, J.S.;
Rinaldi, C. Thermal Decomposition Synthesis of Iron Oxide Nanoparticles with Diminished Magnetic Dead Layer by Controlled
Addition of Oxygen. ACS Nano 2017, 11, 2284–2303. [CrossRef]
97. Lassenberger, A.; Grünewald, T.A.; van Oostrum, P.D.J.; Rennhofer, H.; Amenitsch, H.; Zirbs, R.; Lichtenegger, H.C.; Reimhult, E.
Monodisperse Iron Oxide Nanoparticles by Thermal Decomposition: Elucidating Particle Formation by Second-Resolved in Situ
Small-Angle X-ray Scattering. Chem. Mater. 2017, 29, 4511–4522. [CrossRef]
98. Torres-Gómez, N.; Nava, O.; Argueta-Figueroa, L.; García-Contreras, R.; Baeza-Barrera, A.; Vilchis-Nestor, A.R. Shape tuning of
magnetite nanoparticles obtained by hydrothermal synthesis: Effect of temperature. J. Nanomater. 2019, 1–15. [CrossRef]
99. Ansar, M.Z.; Atiq, S.; Riaz, S.; Naseem, S. Magnetite Nano-crystallites for Anti-cancer Drug Delivery. Mater. Today Proc. 2015,
2, 5410–5414. [CrossRef]
100. Sharafi, Z.; Bakhshi, B.; Javidi, J.; Adrangi, S. Synthesis of Silica-coated Iron Oxide Nanoparticles: Preventing Aggregation without
Using Additives or Seed Pretreatment. Iran. J. Pharm. Res. IJPR 2018, 17, 386–395. [PubMed]
101. Omelyanchik, A.; Salvador, M.; D’orazio, F.; Mameli, V.; Cannas, C.; Fiorani, D.; Musinu, A.; Rivas, M.; Rodionova, V.;
Varvaro, G.; et al. Magnetocrystalline and surface anisotropy in cofe2o4 nanoparticles. Nanomaterials 2020, 10, 1288. [CrossRef]
102. Na, K.-H.; Kim, W.-T.; Park, D.-C.; Shin, H.-G.; Lee, S.-H.; Park, J.; Song, T.-H.; Choi, W.-Y. Fabrication and characterization of the
magnetic ferrite nanofibers by electrospinning process. Thin Solid Film 2018, 660, 358–364. [CrossRef]
103. Rajarao, G.K.; Lakshmanan, R.; Okoli, C.; Boutonnet, M.; Ja, S. Microemulsion prepared magnetic nanoparticles for phosphate
removal: Time efficient studies. J. Environ. Chem. Eng. 2014, 2, 185–189.
104. Kekalo, K.; Koo, K.; Zeitchick, E.; Baker, I. Microemulsion Synthesis of Iron Core/Iron Oxide Shell Magnetic Nanoparticles and
Their Physicochemical Properties. MRS Proc. 2012, 1416, 9–11. [CrossRef] [PubMed]
105. Salvador, M.; Gutiérrez, G.; Noriega, S.; Moyano, A.; Blanco-López, M.C.; Matos, M. Microemulsion Synthesis of Superparamag-
netic Nanoparticles for Bioapplications. Int. J. Mol. Sci. 2021, 22, 427. [CrossRef]
106. Wang, Y.; Nkurikiyimfura, I.; Pan, Z. Sonochemical Synthesis of Magnetic Nanoparticles. Chem. Eng. Commun. 2014, 202, 616–621.
[CrossRef]
107. Fuentes-garc, A.; Alavarse, A.C.; Carolina, A.; Maldonado, M.; Ibarra, M.R.; Fabia, G. Simple Sonochemical Method to Optimize
the Heating E ffi ciency of Magnetic Nanoparticles for Magnetic Fluid Hyperthermia. ACS Omega 2020, 5, 26357–26364. [CrossRef]
108. Holland, H.; Yamaura, M. Synthesis of Magnetite Nanoparticles by Microwave Irradiation and Characterization. In Proceedings
of the Conference: International Latin-American Conference on Powder Technology, Atibaia, Brazil, 8–10 November 2009;
pp. 434–442.
109. Aivazoglou, E.; Metaxa, E.; Hristoforou, E. Microwave-assisted synthesis of iron oxide nanoparticles in biocompatible organic
environment. AIP Adv. 2018, 8, 048201. [CrossRef]
110. Khan, A.A.; Khan, S.; Khan, S.; Rentschler, S.; Laufer, S.; Deigner, H.-P. Biosynthesis of iron oxide magnetic nanoparticles using
clinically isolated Pseudomonas aeruginosa. Sci. Rep. 2021, 11, 1–10. [CrossRef] [PubMed]
111. Elblbesy, M.A.; Madbouly, A.K.; Hamdan, T.A. Bio-synthesis of magnetite nanoparticles by bacteria. Am. J. Nano Res. Appl. 2014,
2, 98–103.
112. Balakrishnan, G.S.; Rajendran, K.; Kalirajan, J. Microbial synthesis of magnetite nanoparticles for arsenic removal. J. Appl. Biol.
Biotechnol. 2020, 8, 70–75.
113. Lina, S.; Tejeda-benitez, L.; Hinestroza, J.; Pati, D.; Herrera, A. Green synthesis of iron oxide nanoparticles using Cymbopogon
citratus extract and sodium carbonate salt: Nanotoxicological considerations for potential environmental applications. Environ.
Nanotechnol. Monit. Manag. 2020, 14, 100377.
114. Kiwumulo, H.F.; Muwonge, H.; Ibingira, C.; Lubwama, M.; Kirabira, J.B.; Ssekitoleko, R.T. Green synthesis and characterization
of iron-oxide nanoparticles using Moringa oleifera: A potential protocol for use in low and middle income countries. BMC Res.
Notes 2022, 15, 1–8. [CrossRef] [PubMed]
115. Bhuiyan, M.S.H.; Miah, M.Y.; Paul, S.C.; Aka, T.D.; Saha, O.; Rahaman, M.M.; Sharif, M.J.I.; Habiba, O.; Ashaduzzaman, M. Green
synthesis of iron oxide nanoparticle using Carica papaya leaf extract: Application for photocatalytic degradation of remazol
yellow RR dye and antibacterial activity. Heliyon 2020, 6, e04603. [CrossRef]
116. Wu, W.; Wu, Z.; Yu, T.; Jiang, C.; Kim, W.-S. Recent progress on magnetic iron oxide nanoparticles: Synthesis, surface functional
strategies and biomedical applications. Sci. Technol. Adv. Mater. 2015, 16, 023501. [CrossRef] [PubMed]
117. Schwaminger, S.; Syhr, C.; Berensmeier, S. Controlled Synthesis of Magnetic Iron Oxide Nanoparticles: Magnetite or Maghemite?
Crystals 2020, 10, 214. [CrossRef]
118. Iconaru, S.L.; Guégan, R.; Popa, C.L.; Motelica-Heino, M.; Ciobanu, C.S.; Predoi, D. Magnetite (Fe3O4) nanoparticles as adsorbents
for As and Cu removal. Appl. Clay Sci. 2016, 134, 128–135. [CrossRef]
119. Klencsár, Z.; Ábrahám, A.; Szabó, L.; Szabó, E.G.; Stichleutner, S.; Kuzmann, E.; Homonnay, Z.; Tolnai, G. The effect of preparation
conditions on magnetite nanoparticles obtained via chemical co-precipitation. Mater. Chem. Phys. 2018, 223, 122–132. [CrossRef]
120. Darwish, M.S.A.; Kim, H.; Lee, H.; Ryu, C.; Yoon, J. Synthesis of Magnetic Ferrite Nanoparticles with High Hyperthermia
Performance via a Controlled Co-Precipitation Method. Nanomaterials 2019, 9, 1176. [CrossRef] [PubMed]
Processes 2022, 10, 2282 23 of 29

121. Maity, D.; Ding, J.; Xue, J.-M. Synthesis Of Magnetite Nanoparticles By Thermal Decomposition: Time, Temperature, Surfactant
And Solvent Effects. Funct. Mater. Lett. 2008, 1, 189–193. [CrossRef]
122. Vangijzegem, T.; Stanicki, D.; Panepinto, A.; Socoliuc, V.; Vekas, L.; Muller, R.N.; Laurent, S. Influence of Experimental Parameters
of a Continuous Flow Process on the Properties of Very Small Iron Oxide Nanoparticles (VSION) Designed for T1-Weighted
Magnetic Resonance Imaging (MRI). Nanomaterials 2020, 10, 757. [CrossRef]
123. Mourdikoudis, S.; Menelaou, M.; Fiuza-Maneiro, N.; Zheng, G.; Wei, S.; Pérez-Juste, J.; Polavarapu, L.; Sofer, Z. Oleic
acid/oleylamine ligand pair: A versatile combination in the synthesis of colloidal nanoparticles. Nanoscale Horiz. 2022, 7, 941–1015.
[CrossRef] [PubMed]
124. Hydrothermal Synthesis Method for Nanoparticle Synthesis—Techinstro. Available online: https://www.techinstro.com/
hydrothermal-synthesis-method-for-nanoparticle-synthesis/ (accessed on 23 February 2022).
125. Gan, Y.X.; Jayatissa, A.H.; Yu, Z.; Chen, X.; Li, M. Hydrothermal Synthesis of Nanomaterials. J. Nanomater. 2020, 2020, 1–3.
[CrossRef]
126. Darr, J.A.; Zhang, J.; Makwana, N.M.; Weng, X. Continuous Hydrothermal Synthesis of Inorganic Nanoparticles: Applications
and Future Directions. Chem. Rev. 2017, 117, 11125–11238. [CrossRef]
127. Hyun, J.; Osman, I.; Saadullah, G. Magnetite Fe3 O4 (111) Surfaces: Impact of Defects on Structure, Stability, and Electronic
Properties. Chem. Mater. 2015, 27, 5856–5867.
128. Richard, B.; Lemyre, J.-L.; Ritcey, A.M. Nanoparticle Size Control in Microemulsion Synthesis. Langmuir 2017, 33, 4748–4757.
[CrossRef]
129. Kimura, K. Magnetic Properties of Magnetite Ultrafine Particles Prepared by W/O Microemulsion Method. Jpn. J. Appl. Phys.
1987, 26, 713.
130. Gautam, R.K.; Chattopadhyaya, M.C. Functionalized Magnetic Nanoparticles: Adsorbents and Applications BT—Nanomaterials
for Wastewater Remediation. In Nanomater. Wastewater Remediat; Elsevier Inc.: Amsterdam, The Netherlands, 2016; pp. 139–159.
131. Singla, R.; Grieser, F.; Ashokkumar, M. Kinetics and Mechanism for the Sonochemical Degradation of a Nonionic Surfactant. J.
Phys. Chem. A 2009, 113, 2865–2872. [CrossRef]
132. Liu, H.; Ji, S.; Yang, H.; Zhang, H.; Tang, M. Ultrasonic-assisted ultra-rapid synthesis of monodisperse meso-SiO2 @Fe3 O4
microspheres with enhanced mesoporous structure. Ultrason. Sonochem. 2014, 21, 505–512. [CrossRef] [PubMed]
133. Perelshtein, I.; Perkas, N.; Gedanken, A. The Sonochemical Functionalization of Textiles; Elsevier Ltd.: Amsterdam, The Netherlands,
2018; pp. 161–198.
134. Choi, J.; Khim, J.; Neppolian, B.; Son, Y. Enhancement of sonochemical oxidation reactions using air sparging in a 36 kHz
sonoreactor. Ultrason. Sonochemistry 2018, 51, 412–418. [CrossRef] [PubMed]
135. Ruan, Q.; Zhu, Y.; Zeng, Y.; Qian, H.; Xiao, J.; Xu, F.; Zhang, L.; Zhao, D. Ultrasonic-Irradiation-Assisted Oriented Assembly of
Ordered Monetite Nanosheets Stacking. J. Phys. Chem. B 2009, 113, 1100–1106. [CrossRef] [PubMed]
136. Chikan, V.; McLaurin, E.J. Rapid Nanoparticle Synthesis by Magnetic and Microwave Heating. Nanomaterials 2016, 6, 85.
[CrossRef]
137. Yang, G.; Park, S.-J. Conventional and Microwave Hydrothermal Synthesis and Application of Functional Materials: A Review.
Materials 2019, 12, 1177. [CrossRef]
138. Kostyukhin, E.; Kustov, L.M. Microwave-assisted synthesis of magnetite nanoparticles possessing superior magnetic properties.
Mendeleev Commun. 2018, 28, 559–561. [CrossRef]
139. Shu, G.; Wang, H.; Zhao, H.-X.; Zhang, X. Microwave-Assisted Synthesis of Black Titanium Monoxide for Synergistic Tumor
Phototherapy. ACS Appl. Mater. Interfaces 2018, 11, 3323–3333. [CrossRef]
140. Strachowski, T.; Grzanka, E.; Mizeracki, J.; Chlanda, A.; Baran, M.; Małek, M.; Niedziałek, M. Microwave-Assisted Hydrothermal
Synthesis of Zinc-Aluminum Spinel ZnAl2 O4 . Materials 2021, 15, 245. [CrossRef]
141. Eugênia, M.; Brollo, F.; Veintemillas-verdaguer, S.; Salván, C.M.; Morales, P. Key Parameters on the Microwave Assisted Synthesis
of Magnetic Nanoparticles for MRI Contrast Agents. Contrast Media Mol. Imaging 2017, 1–13.
142. Kostyukhin, E.M.; Nissenbaum, V.D.; Abkhalimov, E.V.; Kustov, A.L.; Ershov, B.G.; Kustov, L.M. Microwave-Assisted Synthesis
of Water-Dispersible Humate-Coated Magnetite Nanoparticles: Relation of Coating Process Parameters to the Properties of
Nanoparticles. Nanomaterials 2020, 10, 1558. [CrossRef]
143. Schneider, T.; Löwa, A.; Karagiozov, S.; Sprenger, L.; Gutiérrez, L.; Esposito, T.; Marten, G.; Saatchi, K.; Häfeli, U.O. Facile
microwave synthesis of uniform magnetic nanoparticles with minimal sample processing. J. Magn. Magn. Mater. 2017,
421, 283–291. [CrossRef]
144. Fernández-Barahona, I.; Muñoz-Hernando, M.; Herranz, F. Microwave-Driven Synthesis of Iron-Oxide Nanoparticles for
Molecular Imaging. Molecules 2019, 24, 1224. [CrossRef] [PubMed]
145. Chin, S.F.; Azman, A.; Pang, S.C. Size Controlled Synthesis of Starch Nanoparticles by a Microemulsion Method. J. Nanomater.
2014, 2014, 1–7. [CrossRef]
146. Roh, Y.; Liu, S.V.; Li, G.; Huang, H.; Phelps, T.J.; Zhou, J. Isolation and Characterization of Metal-Reducing Thermoanaerobacter
Strains from Deep Subsurface Environments of the Piceance Basin, Colorado. Appl. Environ. Microbiol. 2002, 68, 6013–6020.
[CrossRef]
147. Batool, F.; Iqbal, M.S.; Khan, S.-U.; Khan, J.; Ahmed, B.; Qadir, M.I. Biologically synthesized iron nanoparticles (FeNPs) from
Phoenix dactylifera have anti-bacterial activities. Sci. Rep. 2021, 11, 1–9. [CrossRef]
Processes 2022, 10, 2282 24 of 29

148. Gareev, K.G.; Grouzdev, D.S.; Kharitonskii, P.V.; Kosterov, A.; Koziaeva, V.V.; Sergienko, E.S.; Shevtsov, M.A. Magnetotactic
Bacteria and Magnetosomes: Basic Properties and Applications. Magnetochemistry 2021, 7, 86. [CrossRef]
149. Perotti, G.F.; Da Costa, L.P. Biological Materials. In RSC Nanoscience and Nanotechnology; Royal Society of Chemistry: London, UK,
2021; Volume 2021, pp. 316–332.
150. Vargas, G.; Cypriano, J.; Correa, T.; Leão, P.; Bazylinski, D.A.; Abreu, F. Applications of Magnetotactic Bacteria, Magnetosomes
and Magnetosome Crystals in Biotechnology and Nanotechnology: Mini-Review. Molecules 2018, 23, 2438. [CrossRef]
151. Usov, N.; Gubanova, E. Application of Magnetosomes in Magnetic Hyperthermia. Nanomaterials 2020, 10, 1320. [CrossRef]
152. Baker, I. Magnetic Nanoparticle Synthesisp; Elsevier Ltd.: Amsterdam, The Netherlands, 2018.
153. Yew, Y.P.; Shameli, K.; Miyake, M.; Kuwano, N.; Khairudin, N.B.B.A.; Mohamad, S.E.B.; Lee, K.X. Green Synthesis of Magnetite
(Fe3 O4 ) Nanoparticles Using Seaweed (Kappaphycus alvarezii) Extract. Nanoscale Res. Lett. 2016, 11, 1–7. [CrossRef]
154. Koczkur, K.M.; Mourdikoudis, S.; Polavarapu, L.; Skrabalak, S.E. Polyvinylpyrrolidone (PVP) in nanoparticle synthesis. Dalton
Trans. 2015, 44, 17883–17905. [CrossRef]
155. Makarov, V.V.; Love, A.J.; Sinitsyna, O.V.; Makarova, S.S.; Yaminsky, I.V.; Taliansky, M.E.; Kalinina, N.O. ‘Green’ nanotechnologies:
Synthesis of metal nanoparticles using plants. Acta Nat. 2014, 6, 35–44. [CrossRef]
156. Parajuli, K.; Sah, A.K.; Paudyal, H. Green Synthesis of Magnetite Nanoparticles Using Aqueous Leaves Extracts of Azadirachta
indica and Its Application for the Removal of As(V) from Water. Green Sustain. Chem. 2020, 10, 117–132. [CrossRef]
157. Prasad, C.; Murthy, P.K.; Krishna, R.H.; Rao, R.S.; Suneetha, V.; Venkateswarlu, P. Bio-inspired green synthesis of RGO/Fe3 O4
magnetic nanoparticles using Murrayakoenigii leaves extract and its application for removal of Pb(II) from aqueous solution. J.
Environ. Chem. Eng. 2017, 5, 4374–4380. [CrossRef]
158. Yusefi, M.; Shameli, K.; Yee, O.S.; Teow, S.-Y.; Hedayatnasab, Z.; Jahangirian, H.; Webster, T.J.; Kuča, K. Green Synthesis of Fe3 O4
Nanoparticles Stabilized by a Garcinia mangostana Fruit Peel Extract for Hyperthermia and Anticancer Activities. Int. J. Nanomed.
2021, 16, 2515–2532. [CrossRef] [PubMed]
159. Tyagi, P.K.; Gupta, S.; Tyagi, S.; Kumar, M.; Pandiselvam, R.; Daştan, S.D.; Sharifi-Rad, J.; Gola, D.; Arya, A. Green Synthesis of
Iron Nanoparticles from Spinach Leaf and Banana Peel Aqueous Extracts and Evaluation of Antibacterial Potential. J. Nanomater.
2021, 2021, 1–11. [CrossRef]
160. Nasiri, J.; Rahimi, M.; Hamezadeh, Z.; Motamedi, E.; Naghavi, M.R. Fulfillment of green chemistry for synthesis of silver
nanoparticles using root and leaf extracts of Ferula persica: Solid-state route vs. solution-phase method. J. Clean. Prod. 2018,
192, 514–530. [CrossRef]
161. Pilati, V.; Gomide, G.; Gomes, R.C.; Goya, G.F. Colloidal Stability and Concentration Effects on Nanoparticle Heat Delivery for
Magnetic Fluid Hyperthermia. Langmuir 2021, 37, 1129–1140. [CrossRef]
162. Cortés-Llanos, B.; Ocampo, S.M.; de la Cueva, L.; Calvo, G.F.; Belmonte-Beitia, J.; Pérez, L.; Salas, G.; Ayuso-Sacido, Á. Influence
of Coating and Size of Magnetic Nanoparticles on Cellular Uptake for In Vitro MRI. Nanomaterials 2021, 11, 2888. [CrossRef]
163. Zhang, H.; Hortal, M.; Dobon, A.; Jorda-Beneyto, M.; Bermudez, J.M. Selection of Nanomaterial-Based Active Agents for
Packaging Application: Using Life Cycle Assessment (LCA) as a Tool. Packag. Technol. Sci. 2016, 30, 575–586. [CrossRef]
164. Bobba, S.; Deorsola, F.A.; Blengini, G.A.; Fino, D. LCA of tungsten disulphide (WS 2 ) nano-particles synthesis: State of art and
from-cradle-to-gate LCA. J. Clean. Prod. 2016, 139, 1478–1484. [CrossRef]
165. Zhang, Z.; Guan, Y.; Xia, T.; Du, J.; Li, T.; Sun, Z.; Guo, C. Influence of exposed magnetic nanoparticles and their application in
chemiluminescence immunoassay. Colloids Surf. A Physicochem. Eng. Asp. 2017, 520, 335–342. [CrossRef]
166. Dembski, S.; Schneider, C.; Christ, B.; Retter, M. Core-Shell Nanoparticles and Their Use for In Vitro and In Vivo Diagnos-
tics; Elsevier Ltd.: Amsterdam, The Netherlands, 2018.
167. Ahmadpoor, F.; Masood, A.; Feliu, N.; Parak, W.J.; Shojaosadati, S.A. The Effect of Surface Coating of Iron Oxide Nanoparticles
on Magnetic Resonance Imaging Relaxivity. Front. Nanotechnol. 2021, 3, 1–12. [CrossRef]
168. Wu, K.; Su, D.; Liu, J.; Saha, R.; Wang, J.-P. Magnetic nanoparticles in nanomedicine: A review of recent advances. Nanotechnology
2019, 30, 502003. [CrossRef] [PubMed]
169. Heuer-Jungemann, A.; Feliu, N.; Bakaimi, I.; Hamaly, M.; Alkilany, A.; Chakraborty, I.; Masood, A.; Casula, M.F.; Kostopoulou, A.;
Oh, E.; et al. The Role of Ligands in the Chemical Synthesis and Applications of Inorganic Nanoparticles. Chem. Rev. 2019, 119,
4819–4880. [CrossRef] [PubMed]
170. Tarkistani, M.; Komalla, V.; Kayser, V. Recent Advances in the Use of Iron–Gold Hybrid Nanoparticles for Biomedical Applications.
Nanomaterials 2021, 11, 1227. [CrossRef]
171. Zaloga, J.; Janko, C.; Agarwal, R.; Nowak, J.; Müller, R.; Boccaccini, A.R.; Lee, G.; Odenbach, S.; Lyer, S.; Alexiou, C. Different
Storage Conditions Influence Biocompatibility and Physicochemical Properties of Iron Oxide Nanoparticles. Int. J. Mol. Sci. 2015,
16, 9368–9384. [CrossRef]
172. Widdrat, M.; Kumari, M.; Tompa, É.; Pósfai, M.; Hirt, A.M.; Faivre, D. Keeping Nanoparticles Fully Functional: Long-Term
Storage and Alteration of Magnetite. Chem. Plus Chem. 2014, 79, 1225–1233. [CrossRef]
173. Shubayev, V.I.; Pisanic, T.R.; Jin, S. Magnetic nanoparticles for theragnostics. Adv. Drug Deliv. Rev. 2009, 61, 467–477. [CrossRef]
174. López-Campos, F.; Candini, D.; Carrasco, E.; Francés, M.A.B.; Candini, D. Nanoparticles applied to cancer immunoregulation.
Rep. Pract. Oncol. Radiother. 2019, 24, 47–55. [CrossRef]
175. Mourdikoudis, S.; Kostopoulou, A.; LaGrow, A.P. Magnetic Nanoparticle Composites: Synergistic Effects and Applications. Adv.
Sci. 2021, 8, 1–57. [CrossRef]
Processes 2022, 10, 2282 25 of 29

176. Singh, G.; Rani, S.; Sharma, G.; Kalra, P.; Singh, N.; Verma, V. Coumarin–derived Organosilatranes: Functionalization at magnetic
silica surface and selective recognition of Hg2+ ion. Sens. Actuators B Chem. 2018, 266, 861–872. [CrossRef]
177. Pham, X.-H.; Hahm, E.; Kim, H.-M.; Son, B.S.; Jo, A.; An, J.; Thi, T.A.T.; Nguyen, D.Q.; Jun, B.-H. Silica-Coated Magnetic Iron
Oxide Nanoparticles Grafted onto Graphene Oxide for Protein Isolation. Nanomaterials 2020, 10, 117. [CrossRef]
178. Stöber, W.; Fink, A.; Bohn, E. Controlled growth of monodisperse silica spheres in the micron size range. J. Colloid Interface Sci.
1968, 26, 62–69. [CrossRef]
179. Park, J.C.; Gilbert, D.A.; Liu, K.; Louie, A.Y. Supporting information Microwave enhanced silica encapsulation of magnetic
nanoparticles. J. Mater. Chem. 2012, 22, 8449–8454. [CrossRef]
180. Malhotra, N.; Audira, G.; Chen, J.-R.; Siregar, P.; Hsu, H.-S.; Lee, J.-S.; Ger, T.-R.; Hsiao, C.-D. Surface Modification of Magnetic
Nanoparticles by Carbon-Coating Can Increase Its Biosafety: Evidences from Biochemical and Neurobehavioral Tests in Zebrafish.
Molecules 2020, 25, 2256. [CrossRef]
181. Baykal, A.; Senel, M.; Unal, B.; Karaoğlu, E.; Sözeri, H.; Toprak, M. Acid Functionalized Multiwall Carbon Nanotube/Magnetite
(MWCNT)-COOH/Fe3O4 Hybrid: Synthesis, Characterization and Conductivity Evaluation. J. Inorg. Organomet. Polym. Mater.
2013, 23, 726–735. [CrossRef]
182. Moreno-Bárcenas, A.; Zapata, J.A.A.; Alcalá, M.E.; Ramírez, J.T.; Hernández, A.M.; García-García, A. Evolution of Nanostructured
Carbon Coatings Quality via RT-CVD and Their Tribological Behavior on Nodular Cast Iron. Metals 2022, 12, 517. [CrossRef]
183. Kyesmen, P.I.; Nombona, N.; Diale, M. A Promising Three-Step Heat Treatment Process for Preparing CuO Films for Photocatalytic
Hydrogen Evolution from Water. ACS Omega 2021, 6, 33398–33408. [CrossRef]
184. Chen, Z.; Dai, X.J.; Lamb, P.R.; du Plessis, J.; Leal, D.R.D.C.; Magniez, K.; Fox, B.L.; Wang, X. Coating and Functionalization of
Carbon Fibres Using a Three-Step Plasma Treatment. Plasma Process. Polym. 2013, 10, 1100–1109. [CrossRef]
185. Schwaminger, S.P.; Bauer, D.; Fraga-García, P.; Wagner, F.E.; Berensmeier, S. Oxidation of magnetite nanoparticles: Impact on
surface and crystal properties. Cryst. Eng. Comm. 2017, 19, 246–255. [CrossRef]
186. Sanchez, L.M.; Alvarez, V.A. Advances in Magnetic Noble Metal/Iron-Based Oxide Hybrid Nanoparticles as Biomedical Devices.
Bioengineering 2019, 6, 75. [CrossRef]
187. Ortega, G.; Reguera, E. Biomedical Applications of Magnetite Nanoparticles; Elsevier Inc.: Amsterdam, The Netherlands, 2019.
188. Shiri, M.S.Z.; Henderson, W.; Mucalo, M.R. A Review of The Lesser-Studied Microemulsion-Based Synthesis Methodologies
Used for Preparing Nanoparticle Systems of The Noble Metals, Os, Re, Ir and Rh. Materials 2019, 12, 1896. [CrossRef]
189. Slimani, S.; Concas, G.; Congiu, F.; Barucca, G.; Yaacoub, N.; Talone, A.; Smari, M.; Dhahri, E.; Peddis, D.; Muscas, G. Hybrid Spinel
Iron Oxide Nanoarchitecture Combining Crystalline and Amorphous Parent Material. J. Phys. Chem. C 2021, 125, 10611–10620.
[CrossRef]
190. Mylkie, K.; Nowak, P.; Rybczynski, P.; Ziegler-Borowska, M. Polymer-Coated Magnetite Nanoparticles for Protein Immobilization.
Materials 2021, 14, 248. [CrossRef]
191. Smit, M.; Lutz, M. Polymer-coated magnetic nanoparticles for the efficient capture of Mycobacterium tuberculosis (Mtb). SN Appl.
Sci. 2020, 2, 1–12. [CrossRef]
192. Mirshahghassemi, S.; Cai, B.; Lead, J.R. Evaluation of polymer-coated magnetic nanoparticles for oil separation under environ-
mentally relevant conditions: Effect of ionic strength and natural organic macromolecules. Environ. Sci. Nano 2016, 3, 780–787.
[CrossRef]
193. Kim, D.; Yu, M.K.; Lee, T.S.; Park, J.J.; Jeong, Y.Y.; Jon, S. Amphiphilic polymer-coated hybrid nanoparticles as CT/MRI dual
contrast agents. Nanotechnology 2011, 22, 155101. [CrossRef]
194. Huang, Y.-F.; Liu, Q.-H.; Li, K.; Li, Y.; Chang, N. Magnetic iron(III)-based framework composites for the magnetic solid-phase
extraction of fungicides from environmental water samples. J. Sep. Sci. 2017, 41, 1129–1137. [CrossRef]
195. Sommertune, J.; Sugunan, A.; Ahniyaz, A.; Bejhed, R.S.; Fornara, A. Polymer / Iron Oxide Nanoparticle Composites—A Straight
Forward and Scalable Synthesis Approach. Int. J. Mol. Sci. 2015, 16, 19752–19768. [CrossRef]
196. Li, Y.; Wang, N.; Huang, X.; Li, F.; Davis, T.P.; Qiao, R.; Ling, D. Polymer-Assisted Magnetic Nanoparticle Assemblies for
Biomedical Applications. ACS Appl. Bio Mater. 2019, 3, 121–142. [CrossRef]
197. Beyou, E.; Bourgeat-Lami, E. Organic–inorganic hybrid functional materials by nitroxide-mediated polymerization. Prog. Polym.
Sci. 2021, 121, 101434. [CrossRef]
198. Behrens, S.; Appel, I. Magnetic nanocomposites. Curr. Opin. Biotechnol. 2016, 39, 89–96. [CrossRef]
199. Demirelli, M.; Karaoglu, E.; Baykal, A.; Sozeri, H. M-hexaferrite–APTES/Pd(0) Magnetically Recyclable Nano Catalysts (MRCs).
J. Inorg. Organomet. Polym. Mater. 2013, 23, 1274–1281. [CrossRef]
200. Karaoglu, E.; Baykal, A. CoFe2O4–Pd (0) Nanocomposite: Magnetically Recyclable Catalyst. J. Supercond. Nov. Magn. 2014,
27, 2041–2047. [CrossRef]
201. Junejo, Y.; Baykal, A.; Sözeri, H. Simple hydrothermal synthesis of Fe3 O4 -PEG nanocomposite. Open Chem. 2013, 11, 1527–1532.
[CrossRef]
202. Watt, J.; Collins, A.M.; Vreeland, E.C.; Montaño, G.A.; Huber, D.L. Magnetic Nanocomposites and Their Incorporation into Higher
Order Biosynthetic Functional Architectures. ACS Omega 2018, 3, 503–508. [CrossRef]
203. Alveroǧlu, E.; Sözeri, H.; Baykal, A.; Kurtan, U.; Şenel, M. Fluorescence and magnetic properties of hydrogels containing Fe3 O4
nanoparticles. J. Mol. Struct. 2013, 1037, 361–366. [CrossRef]
Processes 2022, 10, 2282 26 of 29

204. Demir, A.; Baykal, A.; Sözeri, H.; Topkaya, R. Low temperature magnetic investigation of Fe3O4 nanoparticles filled into
multiwalled carbon nanotubes. Synth. Met. 2014, 187, 75–80. [CrossRef]
205. Akal, Z.; Alpsoy, L.; Baykal, A. Biomedical applications of SPION@APTES@PEG-folic acid@carboxylated quercetin nanodrug on
various cancer cells. Appl. Surf. Sci. 2016, 378, 572–581. [CrossRef]
206. Hulla, J.E.; Sahu, S.C.; Hayes, A.W. Nanotechnology: History and future. Hum. Exp. Toxicol. 2015, 34, 1318–1321. [CrossRef]
207. Bayda, S.; Adeel, M.; Tuccinardi, T.; Cordani, M.; Rizzolio, F. The History of Nanoscience and Nanotechnology: From Chemical–
Physical Applications to Nanomedicine. Molecules 2020, 25, 112. [CrossRef] [PubMed]
208. Wei, W.; Zhang, X.; Zhang, S.; Wei, G.; Su, Z. Biomedical and bioactive engineered nanomaterials for targeted tumor photothermal
therapy: A review. Mater. Sci. Eng. C 2019, 104, 109891. [CrossRef]
209. Hose, R.C. Prof. Richard Zsigmondy. Nature 1929, 124, 845–846.
210. Weissig, V.; Pettinger, T.K.; Murdock, N. Nanopharmaceuticals (part 1): Products on the market. Int. J. Nanomed. 2014,
9, 4357–4373. [CrossRef]
211. 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. [CrossRef]
212. Kah, J.; Yeo, E.; He, S.; Engudar, G. Gold Nanorods in Photomedicine in Applications of Nanoscience in Photomedicine; Elsevier Inc.:
Amsterdam, The Netherlands, 2015; pp. 221–248. [CrossRef]
213. Kandasamy, G.; Maity, D. Recent advances in superparamagnetic iron oxide nanoparticles (SPIONs) for in vitro and in vivo
cancer nanotheranostics. Int. J. Pharm. 2015, 496, 191–218. [CrossRef]
214. Vemulkar, T.; Mansell, R.; Petit, D.C.M.C.; Cowburn, R.P.; Lesniak, M.S. Highly tunable perpendicularly magnetized synthetic
antiferromagnets for biotechnology applications. Appl. Phys. Lett. 2015, 107, 012403. [CrossRef]
215. Panahi, H.A.; Alaei, H.S. β-Cyclodextrin/thermosensitive containing polymer brushes grafted onto magnetite nano-particles
for extraction and determination of venlafaxine in biological and pharmaceutical samples. Int. J. Pharm. 2014, 476, 178–184.
[CrossRef]
216. Hu, X.; Wang, Y.; Zhang, L.; Xu, M.; Zhang, J.; Dong, W. Design of a pH-sensitive magnetic composite hydrogel based on salecan
graft copolymer and Fe3O4@SiO2nanoparticles as drug carrier. Int. J. Biol. Macromol. 2018, 107, 1811–1820. [CrossRef]
217. Otero-Lorenzo, R.; Dávila-Ibáñez, A.B.; Comesaña-Hermo, M.; Correa-Duarte, M.A.; Salgueiriño, V. Synergy effects of magnetic
silica nanostructures for drug delivery applications. J. Mater. Chem. B 2014, 2, 2645–2653. [CrossRef]
218. Testa-Anta, M.; Ramos-Docampo, M.A.; Comesaña-Hermo, M.; Rivas-Murias, B.; Salgueiriño, V. Raman spectroscopy to unravel
the magnetic properties of iron oxide nanocrystals for bio-related applications. Nanoscale Adv. 2019, 1, 2086–2103. [CrossRef]
[PubMed]
219. Alphandéry, E. Biodistribution and targeting properties of iron oxide nanoparticles for treatments of cancer and iron anemia
disease. Nanotoxicology 2019, 13, 573–596. [CrossRef] [PubMed]
220. Del Sol-Fernández, S.; Portilla-Tundidor, Y.; Gutiérrez, L.; Odio, O.F.; Reguera, E.; Barber, D.F.; Morales, M.P. Flower-like
Mn-Doped Magnetic Nanoparticles Functionalized with αvβ3-Integrin-Ligand to Efficiently Induce Intracellular Heat after
Alternating Magnetic Field Exposition, Triggering Glioma Cell Death. ACS Appl. Mater. Interfaces 2019, 11, 26648–26663.
[CrossRef]
221. Xu, K.; Yao, H.; Hu, J.; Zhou, J.; Zhou, L.; Wei, S. Pre-drug Self-assembled Nanoparticles: Recovering activity and overcoming
glutathione-associated cell antioxidant resistance against photodynamic therapy. Free Radic. Biol. Med. 2018, 124, 431–446.
[CrossRef]
222. Berry, C.C.; Wells, S.; Charles, S.; Curtis, A.S. Dextran and albumin derivatised iron oxide nanoparticles: Influence on fibroblasts
in vitro. Biomaterials 2003, 24, 4551–4557. [CrossRef]
223. Gupta, A.K.; Curtis, A.S. Lactoferrin and ceruloplasmin derivatized superparamagnetic iron oxide nanoparticles for targeting cell
surface receptors. Biomaterials 2003, 25, 3029–3040. [CrossRef]
224. Pöttler, M.; Fliedner, A.; Schreiber, E.; Janko, C.; Friedrich, R.P.; Bohr, C.; Döllinger, M.; Alexiou, C.; Dürr, S. Impact of
Superparamagnetic Iron Oxide Nanoparticles on Vocal Fold Fibroblasts: Cell Behavior and Cellular Iron Kinetics. Nanoscale Res.
Lett. 2017, 12, 1–9. [CrossRef]
225. Ramchandran, V.; Gernand, J.M. Examining the in vivo pulmonary toxicity of engineered metal oxide nanomaterials using a
genetic algorithm-based dose-response-recovery clustering model. Comput. Toxicol. 2019, 13, 100113. [CrossRef]
226. Sadeghi, L.; Babadi, V.Y.; Espanani, H.R. Toxic effects of the Fe2O3 nanoparticles on the liver and lung tissue. Bratisl. Med. J. 2015,
116, 373–378. [CrossRef]
227. Parivar, K.; Fard, F.M.; Bayat, M.; Alavian, S.M.; Motavaf, M. Evaluation of Iron Oxide Nanoparticles Toxicity on Liver Cells of
BALB/c Rats. Iran. Red Crescent Med. J. 2016, 18, e28939. [CrossRef]
228. Osman, N.M.; Sexton, D.; Saleem, I.Y. Toxicological assessment of nanoparticle interactions with the pulmonary system. Nanotoxi-
cology 2019, 14, 21–58. [CrossRef] [PubMed]
229. Omidkhoda, A.; Mozdarani, H.; Movasaghpoor, A.; Pour Fatholah, A.A. Study of apoptosis in labeled mesenchymal stem cells
with superparamagnetic iron oxide using neutral comet assay. Toxicol. Vitr. 2007, 21, 1191–1196. [CrossRef] [PubMed]
230. Li, X.; Wei, Z.; Lv, H.; Wu, L.; Cui, Y.; Yao, H.; Li, J.; Zhang, H.; Yang, B.; Jiang, J. Iron oxide nanoparticles promote the migration
of mesenchymal stem cells to injury sites. Int. J. Nanomed. 2019, 14, 573–589. [CrossRef]
Processes 2022, 10, 2282 27 of 29

231. Huang, D.-M.; Hsiao, J.-K.; Chen, Y.-C.; Chien, L.-Y.; Yao, M.; Chen, Y.-K.; Ko, B.-S.; Hsu, S.-C.; Tai, L.-A.; Cheng, H.-Y.; et al. The
promotion of human mesenchymal stem cell proliferation by superparamagnetic iron oxide nanoparticles. Biomaterials 2009,
30, 3645–3651. [CrossRef] [PubMed]
232. Balas, M.; Din, I.P.; Hermenean, A.; Cinteza, L.; Dinischiotu, A. Exposure to Iron Oxide Nanoparticles Coated with Phospholipid-
Based Polymeric Micelles Induces Renal Transitory Biochemical and Histopathological Changes in Mice. Materials 2021, 14, 2605.
[CrossRef]
233. Hataminia, F.; Noroozi, Z.; Eslam, H.M. Investigation of iron oxide nanoparticle cytotoxicity in relation to kidney cells: A
mathematical modeling of data mining. Toxicol. Vitr. 2019, 59, 197–203. [CrossRef]
234. Serkova, N.J.; Renner, B.; Larsen, B.A.; Stoldt, C.R.; Hasebroock, K.M.; Bradshaw-Pierce, E.L.; Holers, V.M.; Thurman, J.M. Renal
Inflammation: Targeted Iron Oxide Nanoparticles for Molecular MR Imaging in Mice. Radiology 2010, 255, 517–526. [CrossRef]
235. Zhang, W.; Cao, S.; Liang, S.; Tan, C.H.; Luo, B.; Xu, X.; Saw, P.E. Differently Charged Super-Paramagnetic Iron Oxide Nanoparti-
cles Preferentially Induced M1-Like Phenotype of Macrophages. Front. Bioeng. Biotechnol. 2020, 8, 1–10. [CrossRef]
236. Gu, Z.; Liu, T.; Tang, J.; Yang, Y.; Song, H.; Tuong, Z.K.; Fu, J.; Yu, C. Mechanism of Iron Oxide-Induced Macrophage Activation:
The Impact of Composition and the Underlying Signaling Pathway. J. Am. Chem. Soc. 2019, 141, 6122–6126. [CrossRef]
237. Yarjanli, Z.; Ghaedi, K.; Esmaeili, A.; Rahgozar, S.; Zarrabi, A. Iron oxide nanoparticles may damage to the neural tissue through
iron accumulation, oxidative stress, and protein aggregation. BMC Neurosci. 2017, 18, 1–12. [CrossRef]
238. Yang, Z.; Liu, Z.W.; Allaker, R.P.; Reip, P.; Oxford, J.; Ahmad, Z.; Ren, G. A review of nanoparticle functionality and toxicity on
the central nervous system. Nanotechnol. Brain Future 2013, 7, 313–332.
239. Hajsalimi, G.; Taheri, S.; Shahi, F.; Attar, F.; Ahmadi, H.; Falahati, M. Interaction of iron nanoparticles with nervous system: An
in vitro study. J. Biomol. Struct. Dyn. 2017, 36, 928–937. [CrossRef] [PubMed]
240. Apopa, P.L.; Qian, Y.; Shao, R.; Guo, N.L.; Schwegler-Berry, D.; Pacurari, M.; Porter, D.; Shi, X.; Vallyathan, V.; Castranova, V.; et al.
Iron oxide nanoparticles induce human microvascular endothelial cell permeability through reactive oxygen species production
and microtubule remodeling. Part. Fibre Toxicol. 2009, 6, 1–14. [CrossRef] [PubMed]
241. Duan, J.; Du, J.; Jin, R.; Zhu, W.; Liu, L.; Yang, L.; Li, M.; Gong, Q.; Song, B.; Anderson, J.M.; et al. Iron oxide nanoparticles
promote vascular endothelial cells survival from oxidative stress by enhancement of autophagy. Regen. Biomater. 2019, 6, 221–229.
[CrossRef]
242. Wen, T.; Du, L.; Chen, B.; Yan, D.; Yang, A.; Liu, J.; Gu, N.; Meng, J.; Xu, H. Iron oxide nanoparticles induce reversible endothelial-
to-mesenchymal transition in vascular endothelial cells at acutely non-cytotoxic concentrations. Part. Fibre Toxicol. 2019, 16, 1–13.
[CrossRef]
243. Villanueva, A.; Cañete, M.; Roca, A.G.; Calero, M.; Veintemillas-Verdaguer, S.; Serna, C.J.; Morales, M.D.P.; Miranda, R. The
influence of surface functionalization on the enhanced internalization of magnetic nanoparticles in cancer cells. Nanotechnology
2009, 20, 115103. [CrossRef]
244. Yang, J.-X.; Tang, W.-L.; Wang, X.-X. Superparamagnetic iron oxide nanoparticles may affect endothelial progenitor cell migration
ability and adhesion capacity. Cytotherapy 2010, 12, 251–259. [CrossRef]
245. Cochran, D.B.; Wattamwar, P.P.; Wydra, R.; Hilt, J.Z.; Anderson, K.W.; Eitel, R.E.; Dziubla, T.D. Suppressing iron oxide nanoparticle
toxicity by vascular targeted antioxidant polymer nanoparticles. Biomater. 2013, 34, 9615–9622. [CrossRef]
246. Mahmoudi, M.; Hofmann, H.; Rothen-Rutishauser, B.; Petri-Fink, A. Assessing the In Vitro and In Vivo Toxicity of Superparamag-
netic Iron Oxide Nanoparticles. Chem. Rev. 2011, 112, 2323–2338. [CrossRef]
247. Schimpel, C.; Resch, S.; Flament, G.; Carlander, D.; Vaquero, C.; Bustero, I.; Falk, A. A methodology on how to create a real-life
relevant risk profile for a given nanomaterial. ACS Chem. Health Saf. 2018, 25, 12–23. [CrossRef]
248. Sarpong-Kumankomah, S.; Gibson, M.A.; Gailer, J. Organ damage by toxic metals is critically determined by the bloodstream.
Co-Ord. Chem. Rev. 2018, 374, 376–386. [CrossRef]
249. Krug, H.F. Nanosafety Research-Are We on the Right Track? Angew. Chem. Int. Ed. Engl. 2014, 53, 12304–12319. [CrossRef]
[PubMed]
250. Motayagheni, N. Modified Langendorff technique for mouse heart cannulation: Improved heart quality and decreased risk of
ischemia. MethodsX 2017, 4, 508–512. [CrossRef]
251. Tipton, C.M.; Matthes, R.D.; Tcheng, T.; Dowell, R.T.; Vailas, A.C. The use of the Langendorff preparation to study the bradycardia
of training. Med. Sci. Sport. 1977, 9, 220–230.
252. Bell, R.M.; Mocanu, M.M.; Yellon, D.M. Retrograde heart perfusion: The Langendorff technique of isolated heart perfusion. J. Mol.
Cell Cardiol. 2011, 50, 940–950. [CrossRef]
253. Zimmer, H.-G. The Isolated Perfused Heart and Its Pioneers. Physiology 1998, 13, 203–210. [CrossRef]
254. Stone, V.; Johnston, H.; Schins, R.P.F. Development of in vitro systems for nanotoxicology: Methodological considerations in vitro
methods for nanotoxicology Vicki Stone et al. Crit. Rev. Toxicol. 2009, 39, 613–626. [CrossRef]
255. Erofeev, A.; Gorelkin, P.; Garanina, A.; Alova, A.; Efremova, M.; Vorobyeva, N.; Edwards, C.; Korchev, Y.; Majouga, A. Novel
method for rapid toxicity screening of magnetic nanoparticles. Sci. Rep. 2018, 8, 1–11. [CrossRef]
256. Yuen, H.-W.; Becker, W. Iron Toxicity; StatPearls Publishing: Florida, FL, USA, 2019.
257. Jaishankar, M.; Tseten, T.; Anbalagan, N.; Mathew, B.B.; Beeregowda, K.N. Toxicity, mechanism and health effects of some heavy
metals. Interdiscip. Toxicol. 2014, 7, 60–72. [CrossRef]
258. Wong, V.; Lerner, E. Nitric oxide inhibition strategies. Futur. Sci. OA 2015, 1, 1. [CrossRef]
Processes 2022, 10, 2282 28 of 29

259. Li, Q.; Yon, J.-Y.; Cai, H. Mechanisms and Consequences of eNOS Dysfunction in Hypertension. J. Hypertens. 2015, 33, 1128–1136.
[CrossRef] [PubMed]
260. Zhao, Y.; Vanhoutte, P.M.; Leung, S.W. Vascular nitric oxide: Beyond eNOS. J. Pharmacol. Sci. 2015, 129, 83–94. [CrossRef]
[PubMed]
261. Förstermann, U.; Sessa, W.C. Nitric oxide synthases: Regulation and function. Eur. Heart J. 2012, 33, 829–837. [CrossRef] [PubMed]
262. Skrzypiec-Spring, M.; Grotthus, B.; Szelag, ˛ A.; Schulz, R. Isolated heart perfusion according to Langendorff—Still viable in the
new millennium. J. Pharmacol. Toxicol. Methods 2007, 55, 113–126. [CrossRef] [PubMed]
263. Manuel, A.R.-L.; Martinez-Cuevas, P.P.; Rosas-Hernandez, H.; Oros-Ovalle, C.; Bravo-Sanchez, M.; Martinez-Castañon, G.A.;
Gonzalez, C. Evaluation of vascular tone and cardiac contractility in response to silver nanoparticles, using Langendorff rat heart
preparation. Nanomed. Nanotechnol. Biol. Med. 2017, 13, 1507–1518. [CrossRef]
264. Vargas, J.R.; Harald, O.; Carmen, N.B.; Karen, G. Magnetic nanoparticle behavior evaluation on cardiac tissue contractility
through Langendorff rat heart technique as a nanotoxicology parameter. Appl. Nanosci. 2021, 11, 2383–2396. [CrossRef]
265. Thorat, N.D.; Otari, S.V.; Patil, R.M.; Bohara, R.A.; Yadav, H.M.; Koli, V.B.; Chaurasia, A.K.; Ningthoujam, R.S. Synthesis,
Characterization and Biocompatibility of Chitosan functionalized superparamagnetic nanoparticles for heat activated curing of
cancer cells Published. Dalton Trans. 2014, 43, 17343–17351. [CrossRef]
266. Kumar, P.; Nagarajan, A.; Uchil, P. Analysis of Cell Viability by the Lactate Dehydrogenase Assay. Cold Spring Harb. Protoc. 2018,
2018, 465–469. [CrossRef]
267. Kumar, A.; Gupta, M. Synthesis and surface engineering of iron oxide nanoparticles for biomedical applications. Biomaterials
2005, 26, 3995–4021.
268. Spirou, S.V.; Lima, S.A.C.; Bouziotis, P.; Vranješ-Djurić, S.; Efthimiadou, E.; Laurenzana, A.; Barbosa, A.I.; Garcia-Alonso, I.;
Jones, C.; Jankovic, D.; et al. Recommendations for In Vitro and In Vivo Testing of Magnetic Nanoparticle Hyperthermia
Combined with Radiation Therapy. Nanomaterials 2018, 8, 306. [CrossRef]
269. Jespersen, B.; Tykocki, N.R.; Watts, S.W.; Cobbett, P.J. Measurement of smooth muscle function in the isolated tissue bath-
applications to pharmacology research. J. Vis. Exp. 2015, 95, e52324. [CrossRef] [PubMed]
270. Gonzalez, C.; Corbacho, A.M.; Eiserich, J.P.; Garcia, C.; Lopez-Barrera, F.; Morales-Tlalpan, V.; Barajas-Espinosa, A.;
Diaz-Muñoz, M.; Rubio, R.; Lin, S.-H.; et al. 16K-Prolactin Inhibits Activation of Endothelial Nitric Oxide Synthase, Intracellular
Calcium Mobilization, and Endothelium-Dependent Vasorelaxation. Endocrinology 2004, 145, 5714–5722. [CrossRef] [PubMed]
271. Bachler, G.; von Goetz, N.; Hungerbühler, K. A physiologically based pharmacokinetic model for ionic silver and silver
nanoparticles. Int. J. Nanomed. 2013, 8, 3365–3382.
272. Al-Jamal, K.T.; Bai, J.; Wang, J.T.-W.; Protti, A.; Southern, P.; Bogart, L.; Heidari, H.; Li, X.; Cakebread, A.; Asker, D.; et al. Magnetic
Drug Targeting: Preclinical in Vivo Studies, Mathematical Modeling, and Extrapolation to Humans. Nano Lett. 2016, 16, 5652–5660.
[CrossRef]
273. Corá, L.; Romeiro, F.; Stelzer, M.; Américo, M.; Oliveira, R.; Baffa, O.; Miranda, J. AC biosusceptometry in the study of drug
delivery. Adv. Drug Deliv. Rev. 2005, 57, 1223–1241. [CrossRef]
274. Prospero, A.G.; Fidelis-De-Oliveira, P.; Soares, G.A.; Miranda, M.F.; Pinto, L.A.; Dos Santos, D.C.; Silva, V.D.S.; Zufelato, N.;
Bakuzis, A.F.; Miranda, J.R. AC biosusceptometry and magnetic nanoparticles to assess doxorubicin-induced kidney injury in
rats. Nanomedicine 2020, 15, 511–525. [CrossRef]
275. Alphandéry, E. Bio-synthesized iron oxide nanoparticles for cancer treatment. Int. J. Pharm. 2020, 586, 119472. [CrossRef]
276. Wei, H.; Hu, Y. Superparamagnetic Iron Oxide Nanoparticles: Cytotoxicity, Metabolism, and Cellular Behavior in Biomedicine
Applications. Int. J. Nanomed. 2021, 16, 6097. [CrossRef]
277. Khan, L.U.; Petry, R.; Paula, A.J.; Knobel, M.; Ste, D. Protein Corona Formation on Magnetic Nanoparticles Conjugated with
Luminescent Europium Complexes. ChemNanoMat 2018, 4, 1202–1208. [CrossRef]
278. Nedyalkova, M.; Donkova, B.; Romanova, J.; Tzvetkov, G.; Madurga, S.; Simeonov, V. Iron oxide nanoparticles—In vivo/in vitro
biomedical applications and in silico studies. Adv. Colloid Interface Sci. 2017, 249, 192–212. [CrossRef]
279. Kostal, J. Computational Chemistry in Predictive Toxicology: Status Quo et Quo Vadis? 1st ed.; Elsevier: Amsterdam, The Netherlands,
2016; Volume 10.
280. Antonelli, A.; Sfara, C.; Weber, O.; Pison, U.; Manuali, E.; Salamida, S.; Magnani, M. Characterization of ferucarbotran-loaded
RBCs as long circulating magnetic contrast agents. Nanomedicine 2016, 11, 2781–2795. [CrossRef] [PubMed]
281. European Commission. Nanoreg Data Logging Templates for the Environmental, Health and Safety Assessment of Nanomaterials;
European Commission: Luxembourg, 2013.
282. European Union. Regulation (Eu) No 1169/2011 of the European Parliament; European Union: Luxembourg, 2011.
283. ISO 14577-1:2015; Metallic Materials—Instrumented Indentation Test for Hardness and Materials Parameters–Part 1: Test Method.
ISO: Geneva, Switzerland, 2015. Available online: https://www.iso.org/obp/ui/#iso:std:iso:14577:-1:ed-2:v1:en (accessed on 24
August 2022).
284. National Science and Technology Council Committee on Technology. National Nanotechnology Initiative: Strategic Plan National Science
and Technology Council Subcommittee on Nanoscale Science, Engineering, and Technology Committee on Technology About the National Science
and Technology Council; National Science and Technology Council Committee on Technology: Washington, DC, USA, 2014.
285. FDA. Guidance for Industry Considering Whether an FDA-Regulated Product Involves the Application of Nanotechnology.
Biotechnol. Law Rep. 2011, 30, 613–616. [CrossRef]
Processes 2022, 10, 2282 29 of 29

286. ECHA. Appendix R7-1 for Nanoforms Applicable to Chapter R7a Endpoint Specific Guidance; ECHA: Helsinki, Finland, 2021.
287. Canadian Enviromental Protection Act. Framework for the Risk Assessment of Manufactured Nanomaterials under the Canadian
Environmental Protection Act, 1999 Draft Environment and Climate Change Canada Health Canada Draft June 2022 Executive Summary;
Canadian Enviromental Protection Act: Victoria, BC, Canada, 2022.
288. European Chemicals Agency. Understanding REACH—ECHA. 2018. Available online: https://echa.europa.eu/regulations/
reach/understanding-reach (accessed on 26 September 2018).
289. Sukhanova, A.; Bozrova, S.; Sokolov, P.; Berestovoy, M.; Karaulov, A.; Nabiev, I. Dependence of Nanoparticle Toxicity on Their
Physical and Chemical Properties. Nanoscale Res. Lett. 2018, 13, 44. [CrossRef]
290. Sharma, G.; Kodali, V.; Gaffrey, M.; Wang, W.; Minard, K.R.; Karin, N.J.; Teeguarden, J.G.; Thrall, B.D. Iron oxide nanoparticle
agglomeration influences dose rates and modulates oxidative stress-mediated dose-response profiles in vitro. Nanotoxicology
2014, 8, 663–675. [CrossRef] [PubMed]
291. Gao, X.; Lowry, G.V. Progress towards standardized and validated characterizations for measuring physicochemical properties of
manufactured nanomaterials relevant to nano health and safety risks. NanoImpact 2018, 9, 14–30. [CrossRef]
292. Barik, B.K.; Mishra, M. Nanoparticles as a potential teratogen: A lesson learnt from fruit fly. Nanotoxicology 2018, 13, 258–284.
[CrossRef]
293. Yu, X.; Hong, F.; Zhang, Y.-Q. Bio-effect of nanoparticles in the cardiovascular system. J. Biomed. Mater. Res. Part A 2016,
104, 2881–2897. [CrossRef]
294. Abdelsattar, A.S.; Dawoud, A.; Helal, M.A. Interaction of nanoparticles with biological macromolecules: A review of molecular
docking studies. Nanotoxicology 2020, 15, 66–95. [CrossRef]
295. Simeonidis, K.; Mourdikoudis, S.; Kaprara, E.; Mitrakas, M.; Polavarapu, L. Inorganic engineered nanoparticles in drinking water
treatment: A critical review. Environ. Sci. Water Res. Technol. 2015, 2, 43–70. [CrossRef]

You might also like