Next Article in Journal
Ionic Strength of Methylcellulose-Based Films: An Alternative for Modulating Mechanical Performance and Hydrophobicity for Potential Food Packaging Application
Next Article in Special Issue
Microalgae Polysaccharides: An Alternative Source for Food Production and Sustainable Agriculture
Previous Article in Journal
Modification of Orange Bagasse with Reactive Extrusion to Obtain Cellulose-Based Materials
Previous Article in Special Issue
Biofunctionalization of Endolysins with Oligosacharides: Formulation of Therapeutic Agents to Combat Multi-Resistant Bacteria and Potential Strategies for Their Application
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Biofunctionalized Nanomaterials: Alternative for Encapsulation Process Enhancement

1
Tecnológico Nacional de México/I.T.Tepic, Laboratorio de Investigación en Alimentos, Avenida Tecnológico No. 2595, Tepic 63175, Nayarit, Mexico
2
Laboratorio de Investigación en Materiales, Agua y Energía, Departamento de Ingenierías, Centro Universitario de Los Altos, Universidad de Guadalajara, Av. Rafael Casillas Aceves No. 1200, Tepatitlán de Morelos 47620, Jalisco, Mexico
3
Instituto de Farmacobiología, Universidad de la Cañada, Carretera Teotitlán-San Antonio Nanahuatipán Km 1.7 s/n., Paraje Titlacuatitla, Teotitlán de Flores Magón 68540, Oaxaca, Mexico
*
Authors to whom correspondence should be addressed.
Submission received: 18 February 2022 / Revised: 8 May 2022 / Accepted: 10 May 2022 / Published: 14 May 2022
(This article belongs to the Collection Current Opinion in Polysaccharides)

Abstract

:
In recent years, interest in the development of nanometric materials with specific characteristics has grown; however, there are few scientific contributions that associate encapsulation methodologies and matrices with the particle objective (metabolic directions, type of administration, biological impact, and biocompatibility). This review focuses on describing the benefits and disadvantages of different techniques for designing custom particles and alternatives for the biofunctionalization nanomaterials regarding the biological impact of a nanomaterial with potential use in foods known as nutraceuticals. The study of optical properties, physicochemical factors, and characteristics such as rheological can predict its stability in the application matrix; however, not only should the characterization of a nanocomposite with applications in food be considered, but also the biological impact that it may present.

Graphical Abstract

1. Introduction

In recent years, the impact of nanoscience in research has contributed to the design of nanometric structures of advanced high-performance materials. Given their unique optical, electro-negative, and catalytic properties, some nanoparticles, such as inorganic ones, are attractive as building parts [1], and these characteristics can be modulated by altering the surface state, functionality, and even nanoparticle shape. Other types of structures are made with different polymeric matrices such as polyvinyl alcohol, poly(lactide-co-glycolide), alginic acid, chitosan, etc., which generally have a larger size but have been considered to have the greatest biocompatibility to processes of digestion [2,3]. For the development of nanostructures, simple encapsulation processes are required, such as emulsion and solvent evaporation, or more thorough synthesis processes, for example, sol-gel [4], nanoprecipitation [5], thermal evaporation [6], or those assisted by specific technologies such as spray drying, spray freeze drying [7], solvothermal [8], microwave [9], etc.
An important aspect in the synthesis of new materials is the biofunctionalization, which involves modification of the physicochemical properties of the surface of any material, which allows a wide variety of applications, with a particular biological interest (as an implant, a prosthesis, or any additive administered topically or orally in a living being) to improve the biological response of the organism to the material that has been manufactured [10]. This new concept allows the improvement of surface properties to obtain beneficial effects such as better digestibility and bioaccessibility of encapsulated compounds, as well as faster response time for the biological effect, among other benefits. Therefore, there are alternatives in silico that allow for predicting the compatibility or percentage of charges of different materials for particle formation [11,12].
In this sense, nanoparticles, due to a wide field of application, have been implemented in various industries, such as pharmaceutical, food, and chemical, among others, for the encapsulation of compounds or even microorganisms. However, the analysis of polymer methodologies or matrices is analyzed with respect to the objective of the nanostructure that is intended to be designed [13]. Conversely, it is compelling to characterize the material with certain methods such as dynamic light scattering (DLS), from which the Zeta potential (ZP), polydispersibility index (PdI), and particle size can be measured [14]. DLS can predict the stability in a matrix in different systems, such as a colloid. Fourier-transform infrared spectroscopy (FTIR), Ultraviolet–visible spectroscopy (UV-vis), and Raman spectroscopy provide structure-related information [15], scanning electron microscopy (SEM) shows the morphology of the materials, and with simple uranium staining techniques, the material encapsulated is demonstrated with transmission electron microscopy (TEM) [16].
A relevant factor to evaluate is biocompatibility, for which there are in vitro assays to evaluate the possible damage of nanomaterials in tissues. Among the possible effects that depend on the uptake and internalization of nanoparticles in the cells are short-term cytotoxicity, which includes the affectation of cellular metabolism, alterations in the integrity of the plasma or lysosomal membrane, alterations in the mitochondrial metabolism, reactive oxygen species (ROS) generation, or genotoxicity [17,18]. Nanoparticles could also generate other long-term effects in cells, such as changes in the structure of the cytoskeleton, which alter cell morphology and reduce the proliferative capacity and defects in cell migration or differentiation [17].
This review considers the most important characteristics of the development of biofunctionalized materials, the tools for their characterization, and the biological impact (Figure 1).

2. Biofunctionalization of Nanomaterials Applicable in Food

The implementation of nanoscience methodologies has generated new sources of scientific interest [19]. Specifically, the most important synthesis parameters are the chain length, degree of branching, and crystallinity, particularly in polymers, since these parameters influence the physical, chemical, and transformation behavior. The longer it takes for the polymer chains to be obtained in the synthesis or the higher the degree of branching, the less crystallinity the polymers have, which improves the mechanical resistance, stability, and chemical resistance. Nevertheless, bioactive compounds’ inclusion through different novel methods is involved in the development of biofunctionalized polymeric materials [20]. An example of application is in the development of biopolymers for food packaging. Among the main characteristics that they must meet are degradability, mechanical resistance, and thermal resistance. This can be achieved with the mixture of polymers known as polyvinyl alcohol, crystalline cellulose, and gelatin and allows nanocomposites to be obtained, which are structurally more resistant, with changes that are usually seen in specific FTIR regions [21].
The characteristics and properties of nanomaterials can vary with the parameters and designs used during synthesis. The materials used as the main components of the matrix can be of lipid origin (phospholipids, fatty acids), proteins (albumin, collagen), natural polymers (chitosan, dextran, alginate), semisynthetic (cellulose derivatives, and synthetic polyacrylates (polyacrylamides, polyanhydrides, and polyesters) [22]. These combinations of materials could be of higher molecular weight such as agro-industrial residues; however, the integration of polymeric molecules in the appropriate proportions, considering the structural changes and the interactions between the molecules that make it up, can allow the development of biofunctionalized materials [23].
The chemical structure of the polymer and proteins when used in a suitable encapsulation process (particle size, shape, drug release profile, etc.) can determine the behavior in terms of encapsulation, degradation, and the release of molecules [24]. Due to the wide range of possible combinations of materials, nanoparticles with covers have been classified based on the material from which the core is manufactured and the cover/envelope. Nevertheless, advances in science have implemented other trends to prepare complex materials to mimic the supramolecular architecture of natural structures, with sophisticated and/or multifunctional added capacities (Figure 2). Previous research [25] has shown a proposal of the possible structural conformations for the functionalization of material by incorporating a polymer, a bioactive molecule, and a functionalizing structure, which aims to wrap the structures to confer stability of the particle.

2.1. Bio-Functionalizing Compounds

Bio-functionalization is possible by using compounds of different chemical natures with relevant biological activity, i.e., combining the biological attributes of two or more compounds by enhancing the effect of one of these compounds [26]. One strategy for bio-functionalization involves mixing a compound of low water solubility with one of high water solubility in stoichiometric quantities to ensure the correct interaction between functional groups that can act as anchors [27]. These formulations allow the use of emerging technologies such as an electrospray, where fibers with a particular molecular arrangement can be obtained, which allows for diverse applications, from packaging materials to controlled-release systems for compounds with biological activity [28].
In addition to the above, the properties of some polymeric monomers are exploited through functional groups as anchors for the addition of biological compounds [29]. Biofunctionalizations are performed using both physical and chemical methods of both low-energy (ultrasound, homogenization, emulsification) and high-energy methods (microemulsion polymerization, interfacial polymerization, etc.), always considering the chemical nature of the compounds and their stoichiometric proportions to ensure the expected effect [30].
In biofunctionalization, it is pivotal to consider low-energy interactions such as electrostatic interactions and Van der Waals forces, since these are more likely to be obtained in the combined system. In addition to the participation of a certain surfactant that allows the stability of biofunctionalized compounds in the system of interest, Figure 3 graphically shows one of the ways to consider these energies through structural modeling, which, through a database, allows for predicting the behavior of the structure and translating this into effectiveness or feasibility of using the matrices. Once the biofunctionalization is performed, the system must be maintained with a low humidity concentration to ensure the stability of the components [31]. For this, lyophilization, spray drying, or other physical methods are performed to ensure its subsequent use in vitro and in vitro assays in which the compounds have been considered active [32].
Some biofunctional materials with the structural characteristics and information are provided in Table 1, which are considered important in nanomaterials design since they allow one to predict their behavior. We thereby consider applications together through an analysis of the information collected, which concern parameters, such as solubility and particle size dispersion, among others, that can guide the use of a vehicle such as a patch, drink, or food solid, among others, considering the matrix in which this functionalized material is reported and loaded with a compound that has beneficial characteristics for health [33,34]. The information provided in Table 1 describes the characteristics of the size and composition of the nanoparticles, although the authors do not mention the cytotoxic evaluation.
An important part of biofunctionalization is the risk assessment to obtain these measurements and extract patterns of nanomaterial behavior that can be predicted based on these physicochemical parameters. The toxicity of a specific nanomaterial is correlated with the size of its particles, which can predict the toxicity of a nanomaterial based on a specific parameter in its characterization as FTIR. Through the formation of specific signals in FTIR spectra, some bonds can be identified that can be attributed to possibly toxic substances generated during the synthesis that, through deep analysis, can show a lethal effect on cell lines, but analysis through in vitro tests is also necessary [34].
Table 1. Biofunctionalized materials and relevant information on characterization techniques.
Table 1. Biofunctionalized materials and relevant information on characterization techniques.
MaterialLoadGeneralitiesCharacterizationResultsRef
PLGADonepezil89.67 ± 6.43 nm
Nanoparticles
SEMSpherical nanoparticles[35]
FTIRAt 3008 cm−1 the characteristic peak of aromatic CH group stretch appeared and at 2924 cm−1 the characteristic peak for aliphatic CH2 stretch appeared. Peaks appeared at 1690 cm−1 and 1589 cm−1 corresponding to the C = O carbonyl stretching and aromatic C = C stretches, respectively
XRDPLGA: 20–20° characteristic signal
Load: signals were shown at 5°, 15° y 20°
ChitosanKaempferol137.51–272.91 nm
PdI = 0.25
Nanoparticles
ZPPositive zeta-potential values from + 18.5 to + 38.1 mV.[36]
SEMNanoparticles are spherical in shape, and uniform formation.
FTIRCharacteristic absorption bands: 3324 cm−1 (O─H stretch), 1661 cm−1 (C = O), 1604 cm−1 (C = C), 1378 cm−1 (C─OH), 1257 cm−1(C─O─C) of chitosan and kaempferol shifted to the 1653 cm−1 and the OH stretch of kaempferol completely disappeared at 3363 cm−1.
XRDCrystalline structure at a diffraction angle of 2°θ 10.80°, 12.50°, 15.90°, 23.95°, 24.36°, 27.46°.
PLACurcumin516 and 601 nm
Fibers
FTIRCharacteristic bands corresponding to carbonyl stretching (1750, 1760 cm−1) in PLA and bands corresponding to phenolic, (-OH), C = O and C = C at 1367, 958, 1613, 1505 cm−1[37]
XRDPLA: broad peak at 2θ = 21.84°; load: peaks at 2θ = 11.90°, 14.31°, 17.12°, 17.87°,21.03°, 23.14°, 24.34°, 25.35°, 27.15°and 28.76°.
DSCThe glass transition of the polymer with the load is increased
PLCQuercetinSpheresZPPotential ranged from −120 to 120 mV.[38]
FTIRCharacteristic bans of quercetin as aromatic bending and stretching (1093–1615 cm−1), –OH phenolic bending (1211–1435 cm−1), and –CO stretching (1654 cm−1) and broad phenolic –OH around 3392 cm−1, polymer –CH stretching (2868–2947 cm−1) and –CO stretching (1724 cm−1).
XRDQuercetin: 2θ of 10.80°, 14.16°, 17.23°, 24.33° and 26.97°, but with polymer 21.08°, 23.47° and 26.80°.
Gelatin + ZnOCefazolinDiameter 47.5 nm
PdI: 0.325
Fibers
ZPZeta potential negative −22.67 mV indicated the stable colloidal.[39]
XRD2θ values of 32.42°, 34.51°, 36.78°, 48.1° confirmed the crystalline nature.
FTIRCharacteristic peaks of cefazolin (1761, 1388, 1285 and 1183 cm−1).
PEGNaproxenParticles
Up to 20 µm
SEMExhibited irregularly shaped particles alongside a few spherical
microspheres with a smooth surface
[40]
FTIRAbsorption peaks of 1724 cm−1 and 1681 cm−1 are related to the carboxylic acid bond and benzene ring and peaks of 1155 cm−1 and 1174 cm−1 are correlated to etheric bond.
Ref: Reference; SEM: Scanning electron microscopy; XRD: X-ray diffraction; DSC: Differential scanning calorimetry; ZP: Zeta potential; FTIR: Fourier-transform infrared spectroscopy; PLGA: Poly (lactide-co-glycolide, PLA: Polylactic acid; ZnO: Zinc oxide; PLC: Poly (caprolactone); PEG: Polyethylene glycol; Millivolts: (mV).

2.2. Predictive and Support Tools for Biofunctionalization

Usually, for functionalization, the combination of the compounds is performed via stoichiometric and chemical reactions [31]. However, at present, the implementation of software, specifically in silico methods, allows the establishment of parameters, concentrations, and potential intramolecular interactions for functionalized biomaterial design. For this, it is necessary to carry out the design of the molecules using one of the free or paid platforms, which subsequently requires minimization of energy using an algorithm such as harder fork, beck tree approximations, etc., Subsequently, calculations are carried out to identify the Highest Occupied Molecular Orbital (HOMO)/Lowest Unoccupied Molecular Orbital (LUMO)/Single Occupied Molecular Orbital (SOMO) in addition to the polar and apolar zones of the designed molecules, with which the most stable configuration can be determined in addition to elucidating whether the stability of the compounds or interactions will be strong or not, depending on the distances and the Gibbs-free energy required for its development [41]. Figure 4 shows the interaction between lupeol and mangiferin using the Spartan Student version 8 software as a preliminary stage for the design of nanoparticles obtained in the comprehensive food research laboratory of the Technological Institute of Tepic, to obtain the adequate proportions of both bioactives to be encapsulated with polylactide-co-glycolide (PLGA) (original figure).
The molecules obtained can be used to estimate molecular proportions (stoichiometric quantities). Finally, the appropriate proportions have been established, and functionalization can be carried out with the concentrations closest to reality and with not-so-complex experimental designs (Taguchi, fractional factorial, mix design). The development of biofunctionalized materials is validated with traditional characterization tests, and together with predictions, more satisfactory results can be obtained [42].
Some of the methodologies most associated with functionalization applied in food are those involving the use of polymers due to their effectiveness in encapsulating and biocompatibility, as shown in Table 2. The emulsion and evaporation of solvents, as well as electrospinning, have been some of the most used methodologies in recent years for biofunctionalization since, regularly, they do not form residues that may affect health. However, the stability of the product obtained must be detailed and controlled, so characterization helps to predict its stability in the desired application. It has been observed that zeta potential inferences at 10 mV in aqueous shades can precipitate, and on the other hand, it is important to evaluate the biological activities that are not affected by these encapsulation methods.
In fact, no inorganic materials have been reported for foods and the biofunctionalization of materials since they are toxic; however, some studies have shown that modifying the synthesis processes with less harmful reagents improves this impact [49].

3. Biological Impact of Biofunctionalized Nanomaterials Consumed in Food

The evolution of nanoparticle preparation methods has been characterized by the need to synthesize non-toxic and viable compounds, simplify procedures, and optimally improve performance and efficiency in encapsulation. Different techniques that are simple and safe for the encapsulation of drugs have been mentioned above, such as nanocapsules, nanospheres, nanofibers, etc. A biofunctionalized surface allows a more controlled and directed release. Therefore, nanoparticle applications are widespread, and their development is very promising [50]. Some of the applications associated with nanoparticle shape are identified in Table 3. For human use, the particle size must be greater than 100 nm and frequently implements the use of biocompatible polymers; therefore, the size and shape of the nanomaterial are important parameters to consider in biofunctionalization.

3.1. Enzymes Associated with the In Vitro Study of the Positive Biological Effect of Nanostructures

The areas covered by research on nanoparticles are numerous as they offer better therapeutic options for disorders of public health importance [57]. Some of them have even been used to treat diseases such as cancer. Biomaterials composed of polymers have been widely studied, such as alginate, chitosan, gelatin, hyaluronic acid, polylactic-co-glycolic acid (PLGA), polylactide (PLA), polycaprolactone (PCL), and polyanionic cellulose (PAC). The most commonly used correspond to PLA and PCL [58] due to the bioavailability in the digestion system, improved encapsulation, controlled release, and cytocompatibility and biocompatibility properties, which improve the therapeutic value of the agents encapsulated in colloidal nanoparticles [59]. Studies on the production, characterization, and application of colloidal nanomaterials are rapidly evolving. For nanostructures applied in the oral cavity, it is necessary to evaluate the behavior under different conditions, including factors such as pH, saliva buffer capacity, contact with the mucosa, and dissemination in dental tissues [60].
  • Cyclooxygenase (COX): COX is related to the inflammatory phenomena, and there are currently immunoassay kits that allow the enzyme inhibition to be quantified from a product associated with inflammation (thromboxanes). Therefore, by reducing the activity of the enzyme, the inflammation can be stopped. There are three types of cyclooxygenase: Cyclooxygenase-1 (COX-1), whose function is to regulate the proliferation of normal or neoplastically transformed cells and is found in all tissues, especially in the kidney and gastrointestinal tract, and participates in the production of prostaglandins involved in normal physiological processes such as the protection of the gastric epithelium, maintenance of renal flow, and platelet aggregation. Cyclooxygenase-2 (COX-2) modulates inflammation and prostanoids pathways; therefore, the objective of the design of nanomaterials with anti-inflammatory properties seeks the selective inhibition of COX-2, and current existing immunoassays allow for evaluating this type of effect [59,60].
  • Topoisomerase (TOP): The exacerbated replication of cells, such as the development of carcinoma, is also associated with enzymatic dysregulation, such as TOP, which is the enzyme responsible for maintaining the tertiary DNA structure throughout the cell cycle, being the one in charge of the winding and unwinding of DNA strands during synthesis, replication, condensation, and recondensation. Three types of topoisomerase DNA have been characterized according to their catalytic properties, energy expenditure, and protein structure: Topoisomerase type I DNA (TOP-I) is involved in the opening of the DNA so that the copy of the material is carried out genetically, and the enzyme known as TOP-Ib or TOP-III isogenic of TOP-I is not found in all eukaryotic cells and is responsible for causing a break in one of the chains and forms a phosphodiester bond between the 5 ‘phosphate and the OH of tyrosine. The type II topoisomerase DNAs (TOP-II) are associated with the closure of the same post-process structure and act together to maintain the appropriate level of supercoiling so that biological processes such as cell development are carried out. It is interesting to inhibit these enzymes in processes associated with cancerous processes due to the overexpression of the TOP [61,62].

3.2. Cytotoxicity Associated with Nanomaterials

One of the most used tools for the study of toxicity associated with nanoparticle-based treatments, before their application in vitro, is cell cultures. These have several advantages over in vitro studies: Minor complexity, costs, and ethical issues. Additionally, in vitro assays allow one to control the cellular environment and homogeneity both morphologically and compositionally, allowing a deeper understanding of the biological and biochemicals processes that take place during treatments [63]. The internalization of large particles is favored through phagocytosis processes. Non-specific internalization of particles of ~ 1µm enter via mechanisms of pinocytosis, and particles <100 nm enter the cell via endocytosis, mediated by clathrins or caveolins [64]. Table 4 shows the cell viability of the polymeric nanostructure sizes. It is shown that one of the characteristics of biofunctionalized nanoparticles is that they do not generate an invasive effect or cytotoxicity since these materials lack structures with unpaired electrons or molecules that can break the osmotic balance of a cell such as sodium and zinc, even at sizes below 100 nm.
The cytotoxic effects of a nanoparticle-based treatment may be independent of the cell–particle interaction or can be associated with the process of the internalization and transformation of intracellular nanoparticles. In the first case, this type of cytotoxicity is given by nanoparticles’ direct interaction with the culture medium [72]. Due to this, adding particles to the culture medium can alter pH or osmotic pressure and indirectly cause cell death. To avoid confusion, these possible cytotoxicity results not associated with the cell–nanoparticle interaction were suggested, including in the working protocol the previous incubation of the nanoparticles, its decantation, and resuspension in a fresh culture medium. Besides, it is also recommended to conduct cytotoxicity assays using inhibitors of the endocytosis process or with cell types with very low endocytic capacity (for example, erythrocytes) to evaluate the indirect toxicity induced by nanoparticles in the culture medium [73,74].
Hematoxicity: Hemostasis is the set of mechanisms suitable to make the blood circulate in a liquid state in the blood vessels, while also regulating the process whereby when one of these structures is damaged, the mechanisms of clot formation activate to stop bleeding, as well as the dissolution processes of thrombi (fibrinolysis) [75]. Bearing in mind that the preparation of these nanoparticles has a biomedical end, when administered, they will meet the different components of the blood [76]. Therefore, Table 5 shows some studies that demonstrate the hemotoxic effect of nanomaterials associated with the size. One of the most important barriers to the absorption of nanoparticles in food is blood. In Table 5, all the inorganic nanoparticles show different cellular alterations, given by the energetic composition of the materials since this type of charge affects the erythrocyte membrane, leading to anomalies in health. However, at low concentrations, it is unlikely to show this type of effect [77].

4. Conclusions

The biofunctionalization of nanoparticles is a strategy that has made it possible to enhance the biological activity of different bioactive compounds; however, the physical-chemical characteristics acquired by the nanomaterial during the biofunctionalization process can compromise the biological effect, with the shape, size, and origin of the encapsulating material having the greatest effect on its toxicity. These “toxic attributes” can be identified during characterization and verified by in vitro testing. Undoubtedly, the process of synthesis or development of nanoparticles can lead to the formation of undesirable molecules, therefore evidencing the importance of using biofunctionalizing matrices approved as biocompatible. The challenge for the scientific area is to develop safe nanoparticles, with optimal concentrations of a phytochemical or extract considering characteristics such as the size and shape of the designed material. Additionally, the biological activity of the bioactive compound (autoinflammatory, antioxidant, antiproliferative, etc.) and its route of administration, metabolism, and excretion must be considered since this is an essential point that the physicochemical characteristics can influence.
It is important to carry out the characterization of the biofunctionalized materials before the in vitro biological evaluation since the potential toxicity that it may present can be deduced, through which biomaterials that are less toxic, more biocompatible, and metabolically viable for their degradation could be obtained.

Author Contributions

Conceptualization, F.F.R.-C. and J.A.S.-B.; methodology, M.H.-M.; investigation, F.F.R.-C.; writing—original draft preparation, F.F.R.-C.; writing—review and editing, A.P.-L.; project administration, S.G.S.-A.; funding acquisition, J.A.S.-B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Science and Technology of Mexico (CONACYT) for the grant number 787023 awarded for the development of postgraduate studies.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

The National Council of Science and Technology of Mexico (CONACYT).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Rauwel, P.; Küünal, S.; Ferdov, S.; Rauwel, E. A Review on the Green Synthesis of Silver Nanoparticles and Their Morphologies Studied via TEM. Adv. Mater. Sci. Eng. 2015, 2015, 68274. [Google Scholar] [CrossRef] [Green Version]
  2. Urrejola, M.C.; Soto, L.V.; Zumarán, C.C.; Peñaloza, J.P.; Álvarez, B.; Fuentevilla, I.; Haidar, Z.S. Sistemas de Nanopartículas Poliméricas II: Estructura, Métodos de Elaboración, Características, Propiedades, Biofuncionalización y Tecnologías de Auto-Ensamblaje Capa por Capa (Layer-by-Layer Self-Assembly) Polymeric Nanoparticle Systems: Structure, Elabo. Int. J. Morphol. 2018, 36, 1463–1471. [Google Scholar] [CrossRef] [Green Version]
  3. Casa, D.M.; Carraro, T.C.M.M.; De Camargo, L.E.A.; Dalmolin, L.F.; Khalil, N.M.; Mainardes, R.M. Poly(L-lactide) Nanoparticles Reduce Amphotericin B Cytotoxicity and Maintain Its In Vitro Antifungal Activity. J. Nanosci. Nanotechnol. 2015, 15, 848–854. [Google Scholar] [CrossRef] [PubMed]
  4. Pérez-Larios, A.; Lopez, R.; Hernandez-Gordillo, A.; Tzompantzi, F.; Gómez, R.; Torres-Guerra, L.M. Improved hydrogen production from water splitting using TiO 2-ZnO mixed oxides photocatalysts. Fuel 2012, 100, 139–143. [Google Scholar] [CrossRef]
  5. Sirelkhatim, A.; Mahmud, S.; Seeni, A.; Kaus, N.H.M.; Ann, L.C.; Bakhori, S.K.M.; Hasan, H.; Mohamad, D. Review on Zinc Oxide Nanoparticles: Antibacterial Activity and Toxicity Mechanism. Nano-Micro Lett. 2015, 7, 219–242. [Google Scholar] [CrossRef] [Green Version]
  6. Gaspar, D.; Pimentel, A.C.; Mateus, T.; Leitão, J.P.; Soares, J.; Falcão, B.P.; Araújo, A.; Vicente, A.; Filonovich, S.A.; Águas, H.; et al. Influence of the layer thickness in plasmonic gold nanoparticles produced by thermal evaporation. Sci. Rep. 2013, 3, 1469. [Google Scholar] [CrossRef]
  7. Wang, Y.; Kho, K.; Cheow, W.S.; Hadinoto, K. A comparison between spray drying and spray freeze drying for dry powder inhaler formulation of drug-loaded lipid–polymer hybrid nanoparticles. Int. J. Pharm. 2012, 424, 98–106. [Google Scholar] [CrossRef]
  8. Carpenter, M.K.; Moylan, T.E.; Kukreja, R.S.; Atwan, M.H.; Tessema, M.M. Solvothermal Synthesis of Platinum Alloy Nanoparticles for Oxygen Reduction Electrocatalysis. J. Am. Chem. Soc. 2012, 134, 8535–8542. [Google Scholar] [CrossRef]
  9. Vijayakumar, S.; Ponnalagi, A.K.; Nagamuthu, S.; Muralidharan, G. Microwave assisted synthesis of Co3O4 nanoparticles for high-performance supercapacitors. Electrochim. Acta 2013, 106, 500–505. [Google Scholar] [CrossRef]
  10. Bilek, M. Biofunctionalization of surfaces by energetic ion implantation: Review of progress on applications in implantable biomedical devices and antibody microarrays. Appl. Surf. Sci. 2014, 310, 3–10. [Google Scholar] [CrossRef] [Green Version]
  11. Van Loo, S.; Stoukatch, S.; Axisa, F.; Destiné, J.; Van Overstraeten-Schlogel, N.; Flandre, D.; Lefevre, O.; Mertens, P. Low temperature assembly method of microfluidic bio-molecules detection device. In Proceedings of the 2012 3rd IEEE International Workshop on Low Temperature Bonding for 3D Integration, LTB-3D 2012, Tokyo, Japan, 22–23 May 2012; pp. 181–184. [Google Scholar]
  12. Williams, E.H.; Davydov, A.V.; Oleshko, V.P.; Lin, N.J.; Steffens, K.L.; Manocchi, A.K.; Krylyuk, S.; Rao, M.V.; Schreifels, J.A. Biofunctionalization of Si nanowires using a solution based technique. Nanoepitaxy Mater. Devices IV 2012, 8467, 846702. [Google Scholar] [CrossRef]
  13. Jain, D.; Athawale, R.; Bajaj, A.; Shrikhande, S.; Goel, P.N.; Gude, R.P. Studies on stabilization mechanism and stealth effect of poloxamer 188 onto PLGA nanoparticles. Colloids Surfaces B Biointerfaces 2013, 109, 59–67. [Google Scholar] [CrossRef] [PubMed]
  14. Speed, D.; Westerhoff, P.; Sierra-Alvarez, R.; Draper, R.; Pantano, P.; Aravamudhan, S.; Chen, K.L.; Hristovski, K.; Herckes, P.; Bi, X.; et al. Physical, chemical, and in vitro toxicological characterization of nanoparticles in chemical mechanical planarization suspensions used in the semiconductor industry: Towards environmental health and safety assessments. Environ. Sci. Nano 2015, 2, 227–244. [Google Scholar] [CrossRef]
  15. Madhavi, V. Synthesis and Spectral Characterization of Iron Based Micro and Nanoparticles. Iran. J. Energy Environ. 2013, 4, 385–390. [Google Scholar] [CrossRef]
  16. Karagoz, B.; Esser, L.; Duong, H.T.; Basuki, J.S.; Boyer, C.; Davis, T.P. Polymerization-Induced Self-Assembly (PISA)–control over the morphology of nanoparticles for drug delivery applications. Polym. Chem. 2013, 5, 350–355. [Google Scholar] [CrossRef]
  17. Gonzalez-Moragas, L.; Yu, S.-M.; Benseny-Cases, N.; Stürzenbaum, S.; Roig, A.; Laromaine, A. Toxicogenomics of iron oxide nanoparticles in the nematode C. elegans. Nanotoxicology 2017, 11, 647–657. [Google Scholar] [CrossRef] [Green Version]
  18. Yilmaz, Z.; Dogan, A.L.; Ozdemir, O.; Serper, A. Evaluation of the cytotoxicity of different root canal sealers on L929 cell line by MTT assay. Dent. Mater. J. 2012, 31, 1028–1032. [Google Scholar] [CrossRef] [Green Version]
  19. Ravindran, A.; Chandran, P.; Khan, S.S. Biofunctionalized silver nanoparticles: Advances and prospects. Colloids Surf. B Biointerfaces 2013, 105, 342–352. [Google Scholar] [CrossRef]
  20. Wang, A.; Gu, F.; Zhang, L.; Chan, J.; Radovic-Moreno, A.; Shaikh, M.R.; Farokhzad, O.C. Biofunctionalized targeted nanoparticles for therapeutic applications. Expert Opin. Biol. Ther. 2008, 8, 1063–1070. [Google Scholar] [CrossRef] [Green Version]
  21. Oyeoka, H.C.; Ewulonu, C.M.; Nwuzor, I.C.; Obele, C.M.; Nwabanne, J.T. Packaging and degradability properties of polyvinyl alcohol/gelatin nanocomposite films filled water hyacinth cellulose nanocrystals. J. Bioresour. Bioprod. 2021, 6, 168–185. [Google Scholar] [CrossRef]
  22. Valero-Valdivieso, M.F.; Ortegón, Y.; Uscategui, Y. Biopolímeros: Avances Y Perspectivas Biopolymers: Progress And Prospects. Dyna 2013, 80, 171–180. [Google Scholar]
  23. Azmin, S.N.H.M.; Hayat, N.A.B.M.; Nor, M.S.M. Development and characterization of food packaging bioplastic film from cocoa pod husk cellulose incorporated with sugarcane bagasse fibre. J. Bioresour. Bioprod. 2020, 5, 248–255. [Google Scholar] [CrossRef]
  24. Conde, J.; Dias, J.T.; Grazú, V.; Moros, M.; Baptista, P.V.; de la Fuente, J.M. Revisiting 30 years of biofunctionalization and surface chemistry of inorganic nanoparticles for nanomedicine. Front. Chem. 2014, 2, 1–27. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Brel, V.K.; Artyushin, O.I.; Moiseeva, A.; Sharova, E.V.; Buyanovskaya, A.G.; Nelyubina, Y.V. Functionalization of bioactive substrates with a F5SCH = CH moiety. J. Sulfur Chem. 2019, 41, 29–43. [Google Scholar] [CrossRef]
  26. Vacchi, I.A.; Guo, S.; Raya, J.; Bianco, A.; Menard-Moyon, C. Strategies for the Controlled Covalent Double Functionalization of Graphene Oxide. Chem. Eur. J. 2020, 26, 6591–6598. [Google Scholar] [CrossRef]
  27. Andrade, E.H.A.; Maria das Graças, B.Z.; Maia, J.G.S.; Fabricius, H.; Marx, F. Chemical Characterization of the Fruit of Annona squamosa L. Occurring in the Amazon. J. Food Compos. Anal. 2001, 14, 227–232. [Google Scholar] [CrossRef]
  28. Noel, S.; Liberelle, B.; Robitaille, L.; De Crescenzo, G. Quantification of Primary Amine Groups Available for Subsequent Biofunctionalization of Polymer Surfaces. Bioconj. Chem. 2011, 22, 1690–1699. [Google Scholar] [CrossRef] [Green Version]
  29. Samadarsi, R.; Mishra, D.; Dutta, D. Mangiferin nanoparticles fortified dairy beverage as a low glycemic food product: Its quality attributes and antioxidant properties. Int. J. Food Sci. Technol. 2020, 55, 589–600. [Google Scholar] [CrossRef]
  30. Zarghami, V.; Ghorbani, M.; Bagheri, K.P.; Shokrgozar, M.A. In vitro bactericidal and drug release properties of vancomycin-amino surface functionalized bioactive glass nanoparticles. Mater. Chem. Phys. 2020, 241, 122423. [Google Scholar] [CrossRef]
  31. Brahmachari, G.; Choo, C.; Ambure, P.; Roy, K. In vitro evaluation and in silico screening of synthetic acetylcholinesterase inhibitors bearing functionalized piperidine pharmacophores. Bioorgan. Med. Chem. 2015, 23, 4567–4575. [Google Scholar] [CrossRef]
  32. Vedani, A.; Dobler, M.; Hu, Z.; Smieško, M. OpenVirtualToxLab-A platform for generating and exchanging in silico toxicity data. Toxicol. Lett. 2015, 232, 519–532. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Rekha, R.; Mahboob, S.; Ramya, A.K.; Kerthekeyan, S.; Govindarajan, M.; Al-Ghanim, K.A.; Al-Misned, F.; Ahmed, Z.; Vaseeharan, B. Synthesis and Bio-physical Characterization of Crustin Capped Zinc Oxide Nanoparticles, and Their Photocatalytic, Antibacterial, Antifungal and Antibiofilm Activity. J. Clust. Sci. 2020, 32, 843–855. [Google Scholar] [CrossRef]
  34. Avvakumova, S.; Colombo, M.; Tortora, P.; Prosperi, D. Biotechnological approaches toward nanoparticle biofunctionalization. Trends Biotechnol. 2014, 32, 11–20. [Google Scholar] [CrossRef] [PubMed]
  35. Bhavna; Md, S.; Ali, M.; Baboota, S.; Sahni, J.K.; Bhatnagar, A.; Ali, J. Preparation, characterization, in vivo biodistribution and pharmacokinetic studies of donepezil-loaded PLGA nanoparticles for brain targeting. Drug Dev. Ind. Pharm. 2014, 40, 278–287. [Google Scholar]
  36. Ilk, S.; Sağlam, N.; Özgen, M.; Korkusuz, F. Chitosan nanoparticles enhances the anti-quorum sensing activity of kaempferol. Int. J. Biol. Macromol. 2016, 94, 653–662. [Google Scholar] [CrossRef]
  37. Perumal, G.; Pappuru, S.; Chakraborty, D.; Nandkumar, A.M.; Chand, D.K.; Doble, M. Synthesis and characterization of curcumin loaded PLA—Hyperbranched polyglycerol electrospun blend for wound dressing applications. Mater. Sci. Eng. C 2017, 76, 1196–1204. [Google Scholar] [CrossRef]
  38. Kumar, V.D.; Verma, P.R.P.; Singh, S.K. Development and evaluation of biodegradable polymeric nanoparticles for the effective delivery of quercetin using a quality by design approach. LWT Food Sci. Technol. 2015, 61, 330–338. [Google Scholar] [CrossRef]
  39. Rath, G.; Hussain, T.; Chauhan, G.; Garg, T.; Goyal, A.K. Development and characterization of cefazolin loaded zinc oxide nanoparticles composite gelatin nanofiber mats for postoperative surgical wounds. Mater. Sci. Eng. C 2016, 58, 242–253. [Google Scholar] [CrossRef]
  40. Adibkia, K.; Barzegar-Jalali, M.; Maheri-Esfanjani, H.; Ghanbarzadeh, S.; Shokri, J.; Sabzevari, A.; Javadzadeh, Y. Physicochemical characterization of naproxen solid dispersions prepared via spray drying technology. Powder Technol. 2013, 246, 448–455. [Google Scholar] [CrossRef]
  41. Vedani, A.; Dobler, M.; Smieško, M. VirtualToxLab-A platform for estimating the toxic potential of drugs, chemicals and natural products. Toxicol. Appl. Pharmacol. 2012, 261, 142–153. [Google Scholar] [CrossRef]
  42. Kumari, A.; Yadav, S.K.; Pakade, Y.B.; Singh, B.; Yadav, S.C. Development of biodegradable nanoparticles for delivery of quercetin. Colloids Surf. B Biointerfaces 2010, 80, 184–192. [Google Scholar] [CrossRef] [PubMed]
  43. Mathew, A.; Fukuda, T.; Nagaoka, Y.; Hasumura, T.; Morimoto, H.; Yoshida, Y.; Maekawa, T.; Venugopal, K.; Kumar, D.S. Curcumin Loaded-PLGA Nanoparticles Conjugated with Tet-1 Peptide for Potential Use in Alzheimer’s Disease. PLoS ONE 2012, 7, e32616. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Alamdari, R.F.; Hajimirsadeghi, S.S.; Kohsari, I. Synthesis of silver chromate nanoparticles: Parameter optimization using Taguchi design. Inorg. Mater. 2010, 46, 60–64. [Google Scholar] [CrossRef]
  45. Topuz, F.; Uyar, T. Electrospinning of Cyclodextrin Functional Nanofibers for Drug Delivery Applications. Pharmaceutics 2018, 11, 6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Joseph, J.; Sundar, R.; John, A.; Abraham, A. Phytochemical Incorporated Drug Delivery Scaffolds for Tissue Regeneration. Regen. Eng. Transl. Med. 2018, 4, 167–176. [Google Scholar] [CrossRef]
  47. Pandita, D.; Kumar, S.; Poonia, N.; Lather, V. Solid lipid nanoparticles enhance oral bioavailability of resveratrol, a natural polyphenol. Food Res. Int. 2014, 62, 1165–1174. [Google Scholar] [CrossRef]
  48. Heng, B.C.; Zhao, X.; Xiong, S.; Ng, K.W.; Boey, F.Y.-C.; Loo, J.S.-C. Toxicity of zinc oxide (ZnO) nanoparticles on human bronchial epithelial cells (BEAS-2B) is accentuated by oxidative stress. Food Chem. Toxicol. 2010, 48, 1762–1766. [Google Scholar] [CrossRef]
  49. Di Pasqua, A.J.; Sharma, K.K.; Shi, Y.-L.; Toms, B.B.; Ouellette, W.; Dabrowiak, J.C.; Asefa, T. Cytotoxicity of mesoporous silica nanomaterials. J. Inorg. Biochem. 2008, 102, 1416–1423. [Google Scholar] [CrossRef]
  50. Guterres, S.S.; Alves, M.P.; Pohlmann, A.R. Polymeric Nanoparticles, Nanospheres and Nanocapsules, for Cutaneous Applications. Drug Target Insights 2007, 2, 147–157. [Google Scholar] [CrossRef] [Green Version]
  51. Zhang, P.; Qiao, Z.-A.; Dai, S. Recent advances in carbon nanospheres: Synthetic routes and applications. Chem. Commun. 2015, 51, 9246–9256. [Google Scholar] [CrossRef]
  52. Sanna, V.; Lubinu, G.; Madau, P.; Pala, N.; Nurra, S.; Mariani, A.; Sechi, M. Polymeric Nanoparticles Encapsulating White Tea Extract for Nutraceutical Application. J. Agric. Food Chem. 2015, 63, 2026–2032. [Google Scholar] [CrossRef] [PubMed]
  53. Rezaei, A.; Nasirpour, A.; Fathi, M. Application of Cellulosic Nanofibers in Food Science Using Electrospinning and Its Potential Risk. Compr. Rev. Food Sci. Food Saf. 2015, 14, 269–284. [Google Scholar] [CrossRef] [PubMed]
  54. Li, J.; Xu, W.; Li, D.; Liu, T.; Zhang, Y.S.; Ding, J.; Chen, X. Locally Deployable Nanofiber Patch for Sequential Drug Delivery in Treatment of Primary and Advanced Orthotopic Hepatomas. ACS Nano 2018, 12, 6685–6699. [Google Scholar] [CrossRef] [PubMed]
  55. Zhang, T.; Wang, W.; Zhang, D.; Zhang, X.; Ma, Y.; Zhou, Y.; Qi, L. Biotemplated synthesis of cold nanoparticle-bacteria cellulose nanofiber nanocomposites and their application in biosensing. Adv. Funct. Mater. 2010, 20, 1152–1160. [Google Scholar] [CrossRef]
  56. Grossen, P.; Witzigmann, D.; Sieber, S.; Huwyler, J. PEG-PCL-based nanomedicines: A biodegradable drug delivery system and its application. J. Control. Release 2017, 260, 46–60. [Google Scholar] [CrossRef]
  57. Kumari, A.; Yadav, S.K.; Yadav, S.C. Biodegradable polymeric nanoparticles based drug delivery systems. Colloids Surf B Biointerfaces 2010, 75, 1–18. [Google Scholar] [CrossRef]
  58. Lee, B.K.; Yun, Y.; Park, K. PLA micro- and nano-particles. Adv. Drug Deliv. Rev. 2016, 107, 176–191. [Google Scholar] [CrossRef] [Green Version]
  59. Padovani, G.C.; Feitosa, V.P.; Sauro, S.; Tay, F.R.; Durán, G.; Paula, A.J.; Durán, N. Advances in Dental Materials through Nanotechnology: Facts, Perspectives and Toxicological Aspects. Trends Biotechnol. 2015, 33, 621–636. [Google Scholar] [CrossRef]
  60. Créminon, C.; Habib, A.; Maclouf, J.; Pradelles, P.; Grassi, J.; Frobert, Y. Differential measurement of constitutive (COX-1) and inducible (COX-2) cyclooxygenase expression in human umbilical vein endothelial cells using specific immunometric enzyme immunoassays. Biochim. Et Biophys. Acta Lipids Lipid Metab. 1995, 1254, 341–348. [Google Scholar] [CrossRef]
  61. Moro, M.G.; Oliveira, M.D.D.S.; De Oliveira, L.R.; Teixeira, S.A.; Muscará, M.N.; Spolidorio, L.C.; Holzhausen, M. Effects of Selective Versus Non-Selective COX-2 Inhibition on Experimental Periodontitis. Braz. Dent. J. 2019, 30, 133–138. [Google Scholar] [CrossRef]
  62. Sheng, C.; Miao, Z.; Zhang, W. Topoisomerase I Inhibitors Derived from Natural Products: Structure–Activity Relationships and Antitumor Potency. Stud. Nat. Prod. Chem. 2016, 47, 1–28. [Google Scholar] [CrossRef]
  63. Hevener, K.; Verstak, T.A.; Lutat, K.E.; Riggsbee, D.L.; Mooney, J.W. Recent developments in topoisomerase-targeted cancer chemotherapy. Acta Pharm. Sin. B 2018, 8, 844–861. [Google Scholar] [CrossRef] [PubMed]
  64. Laurent, S.; Mahmoudi, M. Superparamagnetic iron oxide nanoparticles: Promises for diagnosis and treatment of cancer. Int. J. Mol. Epidemiol. Genet. 2011, 2, 367–390. [Google Scholar]
  65. Petros, R.A.; DeSimone, J.M. Strategies in the design of nanoparticles for therapeutic applications. Nat. Rev. Drug Discov. 2010, 9, 615–627. [Google Scholar] [CrossRef] [PubMed]
  66. Auffan, M.; Decome, L.; Rose, J.; Orsiere, T.; De Meo, M.; Briois, V.; Chaneac, C.; Olivi, L.; Berge-Lefranc, J.-L.; Botta, A.; et al. In Vitro Interactions between DMSA-Coated Maghemite Nanoparticles and Human Fibroblasts: A Physicochemical and Cyto-Genotoxical Study. Environ. Sci. Technol. 2006, 40, 4367–4373. [Google Scholar] [CrossRef] [PubMed]
  67. Ruiz-cabello, J. Liver and brain imaging through dimercaptosuccinic acid-coated iron oxide nanoparticles R esearch A rticle. Future Med. 2010, 5, 397–408. [Google Scholar]
  68. Luengo, Y.; Nardecchia, S.; Morales, M.P.; Serrano, M.C. Different cell responses induced by exposure to maghemite nanoparticles. Nanoscale 2013, 5, 11428–11437. [Google Scholar] [CrossRef] [Green Version]
  69. Geppert, M.; Hohnholt, M.C.; Thiel, K.; Nürnberger, S.; Grunwald, I.; Rezwan, K.; Dringen, R. Uptake of dimercaptosuccinate-coated magnetic iron oxide nanoparticles by cultured brain astrocytes. Nanotechnology 2011, 22, 145101. [Google Scholar] [CrossRef]
  70. Gold-Smith, F.; Fernandez, A.; Bishop, K. Mangiferin and Cancer: Mechanisms of Action. Nutrients 2016, 8, 396. [Google Scholar] [CrossRef] [Green Version]
  71. Wang, K.; Körstgens, V.; Yang, D.; Hohn, N.; Roth, S.V.; Müller-Buschbaum, P. Morphology control of low temperature fabricated ZnO nanostructures for transparent active layers in all solid-state dye-sensitized solar cells. J. Mater. Chem. A 2018, 6, 4405–4415. [Google Scholar] [CrossRef]
  72. Fu, Y.-S.; Du, X.-W.; Kulinich, S.A.; Qiu, J.-S.; Qin, W.-J.; Li, R.; Sun, A.J.; Liu, J. Stable Aqueous Dispersion of ZnO Quantum Dots with Strong Blue Emission via Simple Solution Route. J. Am. Chem. Soc. 2007, 129, 16029–16033. [Google Scholar] [CrossRef] [PubMed]
  73. Lundqvist, M.; Stigler, J.; Cedervall, T.; Berggård, T.; Flanagan, M.B.; Lynch, I.; Elia, G.; Dawson, K.A. The Evolution of the Protein Corona around Nanoparticles: A Test Study. ACS Nano 2011, 5, 7503–7509. [Google Scholar] [CrossRef] [PubMed]
  74. Ruiz, A.; Ali, L.M.A.; Cáceres-Vélez, P.R.; Cornudella, R.; Gutiérrez, M.; Moreno, J.A.; Piñol, R.; Palacio, F.; Fascineli, M.L.; de Azevedo, R.B.; et al. Hematotoxicity of magnetite nanoparticles coated with polyethylene glycol: In vitro and in vivo studies. Toxicol. Res. 2015, 4, 1555–1564. [Google Scholar] [CrossRef] [Green Version]
  75. Ghaffar, F.R.A.; Hassouna, I.A.; Ibrahim, H.M.; Elelaimy, I.A.; Abd, H.M. The protective effect of hesperidin or garlic oil against the hemotoxicity of diazinon in male albino rats The protective effect of hesperidin or garlic oil against the hemotoxicity of diazinon in male albino rats. J. Biosci. Appl. Res. 2017, 3, 23–36. [Google Scholar]
  76. Liu, T.; Bai, R.; Zhou, H.; Wang, R.; Liu, J.; Zhao, Y.; Chen, C. The effect of size and surface ligands of iron oxide nanoparticles on blood compatibility. RSC Adv. 2020, 10, 7559–7569. [Google Scholar] [CrossRef]
  77. Sasidharan, A.; Chandran, P.; Monteiro-Riviere, N.A. Biocorona Bound Gold Nanoparticles Augment Their Hematocompatibility Irrespective of Size or Surface Charge. ACS Biomater. Sci. Eng. 2016, 2, 1608–1618. [Google Scholar] [CrossRef]
  78. Dahdouh, F.; Bendjeffal, H.; Nouacer, Z.; Moumene, W.; Zeminour, M.E.-H.; Naous, M.; Djebar, H. Selenium Nanoparticles Attenuate Gentamycin-Induced Nephrotoxicity and Hematotoxicity in Female Swiss Albino Mice. BioNanoScience 2019, 9, 356–364. [Google Scholar] [CrossRef]
  79. Khan, M.S.; Qureshi, N.A.; Jabeen, F.; Shakeel, M.; Asghar, M.S. Assessment of Waterborne Amine-Coated Silver Nanoparticle (Ag-NP)-Induced Toxicity in Labeo rohita by Histological and Hematological Profiles. Biol. Trace Elem. Res. 2017, 182, 130–139. [Google Scholar] [CrossRef]
  80. Asghar, M.S.; Qureshi, N.A.; Jabeen, F.; Khan, M.S.; Shakeel, M.; Chaudhry, A.S. Ameliorative Effects of Selenium in ZnO NP-Induced Oxidative Stress and Hematological Alterations in Catla catla. Biol. Trace Elem. Res. 2018, 186, 279–287. [Google Scholar] [CrossRef] [Green Version]
  81. Mazumdar, H.; Haloi, N.A. study on biosynthesis of iron nanoparticles by Pleurotus sp. Veg. Crops Res. Bull. 2013, 8, 5–19. [Google Scholar]
  82. Yahya, R.A.M.; Azab, A.E.; Shkal, K.E.M.; Jm, J. Hepatotoxicity Induced By Copper Oxide and Zinc Oxide Nanoparticles and Their Mixtures in Male Albino Rats Hepatotoxicity Induced By Copper Oxide and Zinc Oxide Nanoparticles and Their Mixtures in Male Albino Rats. J. Biotechnol. 2019, 3, 1–17. [Google Scholar]
Figure 1. Schematic content.
Figure 1. Schematic content.
Polysaccharides 03 00025 g001
Figure 2. Hypothetical structure of nanoparticles [25].
Figure 2. Hypothetical structure of nanoparticles [25].
Polysaccharides 03 00025 g002
Figure 3. Example of in silico visualization of the biofunctionalized polymer [31].
Figure 3. Example of in silico visualization of the biofunctionalized polymer [31].
Polysaccharides 03 00025 g003
Figure 4. Example of HOMO and LUMO surface in the interaction of two molecules from different chemical characteristics.
Figure 4. Example of HOMO and LUMO surface in the interaction of two molecules from different chemical characteristics.
Polysaccharides 03 00025 g004
Table 2. Biofunctionalized compounds by solvent evaporation method and electrospinning.
Table 2. Biofunctionalized compounds by solvent evaporation method and electrospinning.
Biofunctionalization NanomaterialsSynthesis TypeSize (nm)ResultsRef
PLA + quercetinSolvent evaporation method130Enhancing solubility and stability (ZP 15 mV), spheres by SEM.[43]
PLGA + curcumin150–200Enhancing solubility and increase of antioxidant activity (ABTS).[44]
PLGA + cucurbitacin399 ± 66Promote the encapsulation efficiency until 45% [45]
Cyclodextrin + voriconazoleElectrospinning˂800drug delivery is a dynamic and complex process and can be adapted to meet the needs of the targeted application[46]
Polycarpone + plants phytochemicals˂700Non-toxic formulation of phytochemicals for tissue regeneration and repair.[47]
Lecitin +354 ± 12Improvement of drug loading in the lipid nanoparticles with lecitin.[48]
PLGA: Poly (lactide-co-glycolide, PLA: Polylactic acid, ZP: Zeta potential; mV: Millivolts.
Table 3. Shape and size associated with applications.
Table 3. Shape and size associated with applications.
ShapeSynthesisGeneralitiesApplicationRef
Nanocapsules/nanospheresemulsion and solvent evaporationSize: 200–300 nm.Cutaneous[51]
Hydrothermal carbonization (HTC)Size: 20–50 nm.
monodispersed and mesopores
Electrocatalytic[52]
Nanoprecipitation methodSize: 380.80 ± 37.97
PdI: 0.15 ± 0.06
Nutraceutical[53]
NanofibersElectrospinningDiameter: 400–600 nm.Active and intelligent food packaging products[54]
ElectrospinningDiameter: 139.0 ± 25.6 nm.Nutraceutical patch[55]
CoprecipitationDiameter 100–200 nm.Immobilize enzymes, bioelectroanalysis, and bioelectrocatalysis[56]
Table 4. Cell viability associated with the size of the polymeric nanostructures.
Table 4. Cell viability associated with the size of the polymeric nanostructures.
Cell LineSizeEssaysResultsRef
Fibroblasts humans
Normal
450 ± 20WSTLow effect toxic.[65]
Macrophages murine
RAW264.7
80 ± 18 nmMTTNo significant alterations.[66]
HeLa30 to 70 nmMTTNo significant alterations after 24 h.[67]
Fibroblasts L92949 nmMTTNo toxic seven days after.[68]
Crops primary of Astrocytes60 nmAction enzymatic of LDHNo significant alterations.[69]
Cells
oligodendroglia’s
OLN-93
60 nmAction enzymatic of LDHNo significant alterations.[70]
Stem cells derivate of adipocyte20 nmKit-8 for counting mobilNo significant alterations.[71]
Table 5. Hematoxicity associated with inorganic nanoparticles.
Table 5. Hematoxicity associated with inorganic nanoparticles.
MaterialSizeEffectCite
Gold40 nmPromotes erythrocyte scavenging.[78]
Selenium30 to 100 nmHematological alterations in female mice.[79]
Selenium and zinc oxide36–40 nmAmeliorative effect on hematology and antioxidant systems.[80]
Silver17–22 nmAlterations in the production or synthesis of hematological contents.[81]
Zinc oxide30–50 nmInduced changes in hematological parameters.[82]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Razura-Carmona, F.F.; Perez-Larios, A.; Sáyago-Ayerdi, S.G.; Herrera-Martínez, M.; Sánchez-Burgos, J.A. Biofunctionalized Nanomaterials: Alternative for Encapsulation Process Enhancement. Polysaccharides 2022, 3, 411-425. https://0-doi-org.brum.beds.ac.uk/10.3390/polysaccharides3020025

AMA Style

Razura-Carmona FF, Perez-Larios A, Sáyago-Ayerdi SG, Herrera-Martínez M, Sánchez-Burgos JA. Biofunctionalized Nanomaterials: Alternative for Encapsulation Process Enhancement. Polysaccharides. 2022; 3(2):411-425. https://0-doi-org.brum.beds.ac.uk/10.3390/polysaccharides3020025

Chicago/Turabian Style

Razura-Carmona, Francisco Fabián, Alejandro Perez-Larios, Sonia Guadalupe Sáyago-Ayerdi, Mayra Herrera-Martínez, and Jorge Alberto Sánchez-Burgos. 2022. "Biofunctionalized Nanomaterials: Alternative for Encapsulation Process Enhancement" Polysaccharides 3, no. 2: 411-425. https://0-doi-org.brum.beds.ac.uk/10.3390/polysaccharides3020025

Article Metrics

Back to TopTop