Next Article in Journal
Effect of Ceramic and Dentin Thicknesses and Type of Resin-Based Luting Agents on Intrapulpal Temperature Changes during Luting of Ceramic Inlays
Previous Article in Journal
Factor VII Activating Protease (FSAP) and Its Importance in Hemostasis—Part I: FSAP Structure, Synthesis and Activity Regulation: A Narrative Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Sustainable Secondary-Raw Materials, Natural Substances and Eco-Friendly Nanomaterial-Based Approaches for Improved Surface Performances: An Overview of What They Are and How They Work

by
Silvia Sfameni
1,†,
Giulia Rando
1,2,† and
Maria Rosaria Plutino
1,*
1
Institute for the Study of Nanostructured Materials, ISMN—CNR, Palermo, c/o Department of ChiBioFarAm, University of Messina, 98166 Messina, Italy
2
Department of Chemical, Biological, Pharmaceutical and Environmental Sciences (ChiBioFarAm), University of Messina, 98166 Messina, Italy
*
Author to whom correspondence should be addressed.
These authors contributed equally to the paper.
Int. J. Mol. Sci. 2023, 24(6), 5472; https://doi.org/10.3390/ijms24065472
Submission received: 12 February 2023 / Revised: 3 March 2023 / Accepted: 10 March 2023 / Published: 13 March 2023
(This article belongs to the Special Issue Advanced in Functional Hybrid Nanomaterials)

Abstract

:
To meet modern society’s requirements for sustainability and environmental protection, innovative and smart surface coatings are continually being developed to improve or impart surface functional qualities and protective features. These needs regard numerous different sectors, such as cultural heritage, building, naval, automotive, environmental remediation and textiles. In this regard, researchers and nanotechnology are therefore mostly devoted to the development of new and smart nanostructured finishings and coatings featuring different implemented properties, such as anti-vegetative or antibacterial, hydrophobic, anti-stain, fire retardant, controlled release of drugs, detection of molecules and mechanical resistance. A variety of chemical synthesis techniques are usually employed to obtain novel nanostructured materials based on the use of an appropriate polymeric matrix in combination with either functional doping molecules or blended polymers, as well as multicomponent functional precursors and nanofillers. Further efforts are being made, as described in this review, to carry out green and eco-friendly synthetic protocols, such as sol–gel synthesis, starting from bio-based, natural or waste substances, in order to produce more sustainable (multi)functional hybrid or nanocomposite coatings, with a focus on their life cycle in accordance with the circular economy principles.

1. Introduction

In light of the growing global population and urbanization, the demands of modern civilization include the design of more efficient materials and more resistant surfaces to external agents; in this regard, it appears to be quite necessary to safeguard materials to increase their lifespan by reducing their rate of deterioration [1,2]. The continuous search for suitable mechanical improvement and protective agents in commonly used surfaces and items is driving nanotechnology research to develop new nanostructured coatings for various surfaces featuring enhanced properties [3,4,5]. As a matter of fact, through multidisciplinary feedback, and thanks to the great knowledge of nanotechnology, the design and development of (multi)functional nanohybrids and nanocomposites could greatly improve human daily life and well-being. Particularly, innovative Hybrid Organic-Inorganic Materials (HOIM) or rationally designed functional polymeric blends may be developed and applied in different potential application fields, such as opto-electronic [6], textiles [7], blue growth [8], building [9], mechanical [10], energy conversion [11] and storage [12], sensing [13], environmental remediation [14,15], biomedical [16,17,18] and cultural heritage sectors [19], thanks to their improved features and properties. Quite a lot of advanced nanostructured materials may be employed as surface coatings or in various composite forms for a wide range of applications due to implemented properties (i.e., antifouling [20], flame-retardant [21], drug release [22,23], sensing [24], mechanical resistance [25], and even for the absorption and degradation of environmental pollutants [26,27], as shown in Figure 1). Despite that, several approaches for the development of coatings and finishings still rely on the use of petroleum-based raw materials or doping harmful substances that are hazardous to human health and the environment, from their production cycle until their end-of-life [28]. Nevertheless, in order to achieve eco-friendly alternative functional materials and technologies, it is necessary to focus on developing sustainable manufacturing processes by employing natural, waste or organic resources; only in this way could it be possible to produce coatings that are more sustainable, recyclable, and re-usable, paying attention to their life cycles in agreement with the circular economy principles.
The sol–gel technique is one of these synthetic approaches that has gained more interest lately [29]. Due to its benefits (including low process temperature, high homogeneity of the final products, absence of cytotoxicity and high versatility of the corresponding silane precursors in stable binding of functional molecules or either surface), the sol–gel approach is quite often employed for developing coating featuring stable functionalization as well as improved surface properties of different coated substrates. In this regard, a variety of synthetic pathways may be performed to produce innovative materials with opportunely tunable high molecular homogeneity and superior physical and chemical properties [30]. Sol–gel technology has been successfully applied to obtain different functional properties [7,31,32,33,34]. In particular, this procedure aims to prepare a colloidal solution (sol) in which the monomers are forced to interact with each other through subsequent hydrolysis and condensation reaction steps. These reaction steps lead to the formation of a gel, a solid and continuous lattice submerged in a liquid phase, from which the liquid can then be extracted using thermal approaches or other techniques [35]. More specifically, metallic or metalloid elements are usually employed as precursors in the sol–gel process. Organometallic precursors include silane alkoxides, which are the main fundamental elements of the majority of sol–gel processes. These molecules have a general Si(OR)4 formula, where R is a variable organic functional group depending on the specifically chosen alkoxide [36]. These low-cost precursors enable larger-scale applications in order to use them for procedures with improved reproducibility [37]. The hydrolysis and condensation processes, involving silane alkoxides, keep producing larger 3D frameworks with –Si–O–Si– functionalities, as obtained in both acidic (Figure 2a,b) or alkaline (Figure 2c,d) environments [38]. The flexibility of the lattice decreases as the number of –Si–O–Si– bonds rises, which causes an increase in the lattice viscosity until the latter gels [39]. The first industrial use of sol–gel technology is in thin films and coatings production, by employing different deposition approaches, including spinning [40], dipping [41], spraying [42] and doctor blading [43].
Besides the environmentally friendly sol–gel synthesis, bio-based polymers may also provide a more workable alternative to their fossil-based counterparts [44]. According to a recent market analysis by European Bioplastics, the production of bioplastics would increase globally by more than 35% between 2020 and 2025 [45]. Moreover, biomass materials have been employed as structural or common materials since ancient times, as their primary constituents were natural polymers, including lignin, cellulose, starch, chitin, and protein, which ensured sufficient strength and stability [46]. Furthermore, the large subset of naturally occurring materials, coming from agro-industrial wastes or via industrial processing of animals and agricultural products, and consisting of around five billion metric tons annually, can therefore represent a source of precious secondary raw materials for value-added and sustainable applications. For example, lignin, i.e., the second-most common biopolymer found in plant cell walls, is one of the many different types of agricultural biomasses. Large amounts of lignin are produced as a by-product of the pulp and paper industries. Environmental concerns have recently intensified efforts to employ lignin derivatives in value-added applications such as for antifouling coating preparation [47]. These eco-friendly and sustainable solutions and resources are of great interest in the field of nanotechnology. In fact, the green formulations that may be generated from these materials can be modified by adding opportune functional nanofillers, such as nanoparticles, to obtain useful protective and multifunctional coatings for various surfaces and substrates [48].
In light of these factors, the goal of this review is therefore to highlight the recent approaches for the development and characterization of hybrid materials, nanocomposites and functional coatings, also obtained through the employment of different secondary-raw, natural and eco-friendly nanostructured materials for smart, innovative and sustainable applications. In particular, coatings based on functional nanoparticles, molecules, nanofillers, sustainable recycled materials and natural substances, as dispersed appropriately in polymeric matrixes or in combination with proper sol–gel based cross-linkers and dopant agents, will be compared in order to give an overview on the recent sustainable advancement in improved surface performances materials. Furthermore, the molecular and chemical aspects behind the mechanism of action of various functional coatings are discussed to contribute to the rational safe-by-design of enhanced, green, and smart surfaces.

2. Functional Protective Sustainable Coatings

Protective coatings are crucial in today’s society since they may significantly improve our life, comfort, and welfare [49]. As a matter of fact, all surfaces exposed to atmospheric and environmental conditions undergo various forms of deterioration, which are influenced by a variety of causes and agents, the most notable of which are:
  • Material composition and porosity;
  • Accidents, alterations, and previous restorations or treatments;
  • Type of exposure (contaminants, climatic conditions, anthropological activities);
  • Colonization via biological or microbiological agents.
By decreasing or counteracting the deterioration kinetics, it is possible to increase the lifespan and durability of surfaces and materials by acting on the surrounding environment and defending the surface of the material itself [50,51,52].
One of the main agents from which surfaces need to be protected is water; it is the main cause of degradation, leading to a direct or indirect deterioration of materials, also due to the presence of potentially harmful solutes that can accelerate or promote this process [53]. Therefore, an ideal protective coating should exhibit different qualities like:
  • Hydrophobicity;
  • Transpiration capabilities towards water vapors;
  • UV-resistance;
  • Chemical inertness;
  • Strong resistance to harmful chemical agents;
  • Lead to minimal alteration of the appearance of the coated product;
  • Resistance to bacterial and mold colonization.
Buildings, automobiles, textiles, industrial machinery and other infrastructure are constantly preserved to have longer life and durability due to the application of top-coatings, primers, sealants, varnishes and paints. Many of these actual surface protection treatments and coatings may still be environmentally hazardous [54,55,56]. Therefore, the main approach remains the development of unique, high-performing, ecologically friendly and functional coatings starting from bio-based formulations, waste, natural materials and compounds, as well as more sustainable processes. Some of the recent examples and approaches described and discussed in this review are collected in Table 1, reported together with a comparison of the functional agents, deposition approach, coated surfaces and implemented achieved features.

2.1. Anticorrosive Hybrid Nanostructured Coatings Doped with Green Corrosion Inhibitors

Every structure that is subjected to corrosion agents must be protected in order to withstand corrosive forces for the duration of the structure’s required life. Corrosion can be decreased or prevented by using specialized procedures that delay or inhibit anodic or cathodic reactions, or by eliminating conductivity between anodic and cathodic sites. Corrosion is the collective term for all chemical and electrochemical processes that return metals to their lower Gibbs free energy states. This phenomenon is largely caused by electrochemical processes, in which electrons move between two half-cell reactions [78,79,80].
Anodic or metal oxidation processes are one of the half-cell reactions that generate electrons, as shown in Equation (1), in which M represents the metal.
M0 → Mn+ + ne
The produced electrons are involved in the cathodic or reduction reaction, which is the other half of the cell reaction. The pH and the accessibility of different components affect cathodic reactions. Under acidic circumstances, hydrogen and oxygen reduction are the most frequent cathodic reactions, as shown in subsequent Equations (2) and (3):
2H+ + 2e → H2
O2 + 4H+ + 4e → 2H2O
Additionally, water or oxygen reduction processes occur in neutral or alkaline conditions, as shown in the following Equations (4) and (5).
2H2O + 2e → H2 + 2OH
O2 + 2H2O + 4e → 4OH
The procedure for protecting the metal surface often involves two (pre)treatments that are performed in sequence, namely the preparation pretreatment of the target surface, which removes any impurities or remnants of earlier corrosive processes from the metal or substrate to be protected, and the final application of the protective top-coat [81,82]. Actual anticorrosive protective solutions require the employment of hazardous phosphating and chromating procedures, which, despite their efficacy, have considerable and negative environmental and energy consequences [83]. In particular, toxic phosphates and chromium compounds, in fact, lead to the production of large amounts of hazardous sludge that must be disposed. In this regard, new developments in the field of environmentally friendly polymer-based anticorrosive coatings of parts or specific objects/substrates have been studied and described in several research products.
These cutting-edge and environmentally friendly materials/technologies, as depicted in Figure 3, can be divided into five groups: hyperbranched/hybrid polymer technologies, bio-based materials, green corrosion inhibitors, and (super)hydrophobic coatings.
The use of natural additives to enhance the properties of an anticorrosive coating is a new finding that has fascinated researchers as well as product end-users [84].
Carbon-based nanomaterials, such as graphene or graphene oxide, are of great interest due to their powerful anticorrosion feature and barrier properties, which prevent corrosive substances from entering and supplying an electron flow channel between the sacrificial anode and the substrate [86,87].
Moreover, hyperbranched polymers have most recently drawn a lot of interest in the field of polymer research due to their remarkable qualities, including high cross-linking density, superior solubility and high reactivity, thus leading to the further production of low solvent chemical-resistant and long-life coatings [88,89,90].
The application of plant extracts as environmentally benign corrosion inhibitors, such as polyphenols, has also attracted the scientific community’s interest [91]. Its inhibitory impact is due to a unique antioxidant property that causes complexes to form on the metal surface, promoting the formation of a strong barrier that protects the surface from external agents. Various studies have proved the use of pure tannins and other polyphenols as corrosion inhibitors [92,93,94].
To combine the protective feature of graphene oxide and reduce its agglomeration due to π–π stacking, which makes it hard to disperse uniformly in a polymer matrix, different covalent and/or noncovalent functionalization approaches can be employed. In fact, thanks to its organic functionalities (i.e., hydroxyl, carboxyl, epoxy groups), it can react easily to form stable dispersions. An example is represented by the functionalization of graphene oxide with phytic acid, as shown in Figure 4a,b. This latter is a naturally abundant organic coordination molecule with the chemical structure of C6H18O24P6 in which each phosphate group, characterizing the molecule, is connected to a carbon atom of the cyclo-hexamehexol ring [95]. Thanks to its non-toxicity, biocompatibility and environmental friendliness, it has been widely employed in a variety of applications, i.e., food and cosmetic additives, cleaning agents, and complexation of pollutants [96,97,98]. An embedding waterborne host sol–gel matrix can be easily obtained by the use of proper silane precursors and cross-linkers, in order to produce a uniform, homogeneous, durable, stable and functional coating for metal substrates (Figure 4c,d).
In the described study, (3-glycidyloxypropyl)trimethoxysilane, hereafter indicated as GPTMS, and 3-(aminopropyl)triethoxysilane, indicated as APTES, were employed for this purpose.
The results of polarization measurements and a neutral salt spray test confirmed the anticorrosive behavior on steel and aluminum substrates of the final obtained sol–gel based nanostructured hybrid coating doped with phytic acid intercalated graphene oxide [57].
Electrochemical coatings are a popular method for protecting metallic surfaces. These coatings can be made of simple metals as well as binary and ternary alloys, which improve corrosion, oxidation, and wear resistance. An electrochemical deposition is a low-cost option for thin-film synthesis since it does not require complicated or expensive equipment and employs common raw materials. By altering deposition settings, electrodeposition enables the simple control of the coating’s chemical composition, shape, and thickness [99].
Because of its recognized anticorrosion potential and environmental friendliness, the electrodeposition of polyaniline (PANI) coatings onto metal surfaces has gained attention. The anticorrosion effect of such PANI coatings is thought to be accomplished through a variety of mechanisms, including barrier protection, corrosion inhibition, anodic protection, a shift of electrochemical reactions from the metal/coating interface to the coating/electrolyte interface, passivation, and so on. By using a potentiostatic technique, a polyaniline-reduced graphene oxide composite (PANI-rGO) film was electrochemically deposited on 5083 Al alloy. The results revealed that the presence of rGO promoted PANI electropolymerization and improved Al alloy passivation [58].
As mentioned before, tannins are well-known polyphenols coming mainly from pine bark and acacia, featuring anticorrosion inhibition properties. The presence of OH groups in the ortho position of the aromatic ring of tannins leads to their reaction with iron, iron salts and oxidized steel substrates, promoting the formation of mono- and bi-ferric tannate species (tannin-Fe complexes), inhibiting the activity of the corrosion agents. In this regard, a hybrid self-healing coating, containing silane and tannins functionalized zinc oxide nanoparticles, was produced to protect steel substrates. In particular, ZnO nanoparticles, obtained by the method of arc discharge in a controlled atmosphere, were functionalized with APTES, mixed with Pinus radiata-derived tannin and an epoxy resin to obtain the functional hybrid epoxy. Subsequently, the obtained formulation was deposited on ASTM A36 steel plates through a spray coating approach. The anticorrosive properties of the functionalized nanoparticles were demonstrated using electrochemical analysis. Moreover, the contact angle and Kelvin probe delamination studies showed the self-healing capabilities of the film with the substrate, as shown in Figure 5 [59].
Corrosion is a more aggressive-related phenomenon affecting all the surfaces exposed in a marine environment. In this case, the proposed corrosion mechanism for steel surfaces, for example, is as follows (Equations (6)–(9)) [100]:
Fe → Fe2+/3+ + 2e/3e (anode)
O2 + 2H2O + 4e → 4OH (cathode)
Fe2+/Fe3+ + 2Cl/3Cl → FeCl2/FeCl3
FeCl2/FeCl3 + 2H2O/3H2O → Fe(OH)2/Fe(OH)3 + 2HCl/3HCl (unstable)
For these purposes, natural inhibitors for corrosion were also evaluated. An extract of tropical plant Mangifera indica L. leaf (MIL), containing a high quantity of polyphenols, was used to obtain an amorphous silica-based hybrid material, subsequently incorporated in an epoxy coating. In detail, a precipitated amorphous silica functionalized with MIL was dispersed in an epoxy resin and its polyamide curing agent with a ratio of 2:1. Finally, by a dip-coating approach, the dispersion was employed for the treatment of commercial steel substrates (Figure 6a).
The coating developed featured a superior corrosion inhibition performance of 99%, according to electrochemical measurements. The mechanism of action is explained by the formation of a FeO(OH) passive layer and an insoluble stable organometallic complex due to the hydroxylic and carboxylic functionalities of polyphenols in MIL extract preventing the Cl attack, as shown in Figure 6b and described in Equation (10).
Fe–OHads → FeO(OH)ads → Fe–MILads + 2H2O
The results of this study thus showed a unique, environmentally friendly, effective corrosion inhibitor for the defense of steel in saline media [60].

2.2. Eco-Friendly Approaches for Flame-Retardant Coatings Preparation

The majority of polymeric materials are characterized by highly combustible chemical and organic components. However, the polymers’ inherent flammability, resulting from their chemical structures and organic components (Figure 7), limits their practical usage. The most common flame-retardants employed in various industries are mainly based on inorganic, halogenated species and phosphorus-based agents together with nitrogenous compounds (due to their synergistic effect) [101].
Flame retardants are essentially based on the trapping of free radical species involved in (H and OH) oxidation reactions (gas phase) to inhibit them. Even though there are many radical reactions occurring during gas phase combustion, only the following two steps, as shown in Equations (11) and (12), are accountable for the rapid process and the energy released [103]:
H + O2 → OH + O (propagation stage)
OH + CO → CO2 + H (exothermic reaction)
In this regard, researchers are working on the development of new flame-retardant agents and nanocomposites, employing different nanomaterials and compounds such as the ones shown in Figure 8a. The addition of a relatively modest loading level of around 5 wt% of nanomaterials to the final polymeric nanocomposites can reduce their flammability, resulting in a decrease in the heat release rate (HRR) and mass loss rate (MLR) [102]. Despite nanocomposites, naturally derived compounds can also be employed for the development of sustainable flame retardants (Figure 8b).
As a result, it is important to properly rationalize a flame-retardant finishing and consider both its physical and chemical methods of action [102,105]. Several of them, which will be further reviewed, deal with:
  • Encouraging endothermic processes;
  • Creating inert gases that reduce/dilute the air oxygen’s content;
  • Producing an impermeable layer of protection;
  • Introducing flame retardants that can scavenge and eliminate active radicals in the gas and condensed phases;
  • Dehydrating, cyclizing, and cross-linking flame retardants and/or the polymer matrix to create a carbonaceous protective layer.
Besides the sol–gel synthetic method [106,107], a better eco-friendly approach is represented by the use of natural or waste materials, such as coffee biowastes, as sustainable flame retardants. In particular, spent coffee grounds were chemically modified with phosphorus (dimethyl phosphite) in different ratios to obtain the P-coffee derivative. An epoxy resin containing 30 wt% P-coffee demonstrated a considerable reduction in the pHRR (40%) value by flammability analysis in pyrolysis combustion and flow calorimeter analysis. The presence of a carbon source, coffee biowaste, and phosphorus in the epoxy resin aided and sped up the production of a carbonaceous residue (char layer) that served as a barrier against heat and mass diffusion of gases into the gas phase. In the interim, the released phosphorus compounds stopped the spread of the flame by capturing free radicals in the gas phase, as shown in Figure 9 [61].
Through the in-situ synthesis of hydroxyapatite (a bio-derived polycrystalline calcium phosphate ceramic with a hexagonal structure) in the presence of lignocellulose, a novel bio-based flame retardant for poly(lactic acid) (PLA) was developed. In particular, the hydroxyapatite-modified lignocellulose was produced from a milled bagasse-derived lignocellulosic biomass and perylene dianhydride as the grafting agent. The resulting hybrid material incorporates phosphorus, hydroxyl, and aromatic functionalities to work as a long-lasting flame-retardant system. Finally, it was combined with ammonium polyphosphate and integrated into a PLA matrix to achieve higher flame retardancy than pristine PLA [108].
Among the class of cellulosic-derived products, ammonium starch phosphate carbamates (mixed starch esters primarily composed of covalently bound ammonium phosphate groups with trace quantities of carbamate groups) can be applied as new and sustainable coatings with flame-retardancy features. They are produced significantly more easily than cellulose phosphates by adopting a solvent-free eco-friendly method to combine starch with urea (as an “esterification promoter”) and phosphoric acid. Ammonia was found to be a result of ammonium starch phosphate carbamates decomposition and, as inert gas, aids in flame extinguishment [109].
Moreover, other natural derivatives and bio-macromolecules, such as proteins and casein, a phosphorylated protein derived from renewable precursors, find application as green potential flame inhibitors for textile materials due to their capacity to induce the dehydration process of cellulose chains creating a char layer rather than the depolymerization process and their ability to achieve outstanding flame retardancy action through the heat consumption process [110,111]. Additionally, precious inorganic clay compounds, such as halloysite nanotubes (HNTs) characterized by hollow nanotubular structures, are still currently used as a useful synthon for creating polymeric nanocomposite materials with implemented features due to their large surface area and low cost [112,113,114]. Different techniques have been utilized to obtain HNT nanocomposites, among which the melt mixing process seems to be the most popular. To obtain the good dispersion of HNTs into polymeric blends, the HNTs must be functionalized. The inclusion of HNTs enhanced the mechanical, thermal, and flammability qualities of the functionalized surfaces and materials [115,116].
To combine the features of casein and HNTs, a new one-dimensional nanocomposite coating was developed for different textile supports. In particular, a different number of mass loadings of HNTs (10, 30 and 50 wt%) was uniformly dispersed in a rennet casein (RC) solution prepared from renewable skim milk with the help of ultrasonication. The created bio-inspired nanocomposite was then employed to cover various textiles (cotton, polyester and blend of cotton and polyester) with the dry-pad-cure process. The coated textile fabrics’ flammability, toxic gas suppressing, reinforcing, antibacterial, and antiviral characteristics were all greatly improved. In a horizontal test, treated textile fabrics’ flame retardancy reached a rate of burning of zero, compared to 119 mm/min for untreated cotton fabrics. Further, a high LOI value of 35%, as opposed to 18.5% for a blank sample, supports the flame retardancy properties of the green coating developed [62].

2.3. Hydrophobic and Water-Repellent Sustainable Coatings

The use of hydrophobic surfaces in both everyday life and some industrial processes has generated a lot of interest in recent years [117].
Surface composition, as well as its texturing and roughness, have been shown to have a significant impact on its hydrophobicity. In this regard, numerous techniques, such as the deposition of layers made of nonrenewable materials (such as fluorine or hydrocarbon compounds, some wax or organic and inorganic materials), have been opportunely formulated to lower the surface energy and, therefore, improve the water contact angle (WCA) of surfaces [118,119]. In the same way, surfaces with a hierarchical structure consisting of nanostructured texturing were successful in achieving high levels of hydrophobicity and superhydrophobicity [120].
Recent attention has been focused on the design of bio-inspired surfaces by the superhydrophobicity natural model, i.e., the lotus leaf [121]. The majority of hydrophobic surfaces are produced by treatments with fluorine compounds, which have negative environmental effects due to their bioaccumulation [122]. However, the majority of the approaches that have been documented up to this point involve high costs, difficult processes, and the use of hazardous chemicals and solvents. To increase the range of industrial applications for hydrophobic surfaces with controllable morphology, it is still difficult to develop straightforward, quick, inexpensive, and environmentally responsible methods [123].
In the next paragraphs, an overview of the recent advancements in organic–inorganic hybrid and sustainable coatings is reported for the development of nanostructured finishings for cultural heritages and textile surfaces to achieve different improved features.

2.3.1. Cultural Heritages Stones Protection

The development of protective coatings for both transportable and immovable cultural heritage items has drawn increasingly more attention in recent years [124]. In particular, it has mostly been caused by an increased awareness of the need to preserve cultural artifacts and monuments affected by weathering exposure and the existence of reactive airborne chemicals that may interact with the materials and compromise them (Figure 10a) [125]. Several research studies have advanced significantly from the end-of-the-last-century acrylic resins to the modern biomaterials and nanoparticles used today [19,126].
The widely accepted standards of cultural heritage restoration and preservation should offer the qualities of the ideal protective coatings (including transparency, reversibility, compatibility with the surface, long lifetime, ease of synthesis, low-cost maintenance and non-toxicity) and must be typically designed for cultural objects [19].
Synthetic polymers such poly acrylates, siloxanes, and fluorinated polymers have been widely used as protective coatings for stone surfaces in cultural heritages leading to different implemented and protective features (Figure 10b).
Despite their excellent water repellent and optical clarity properties, the continuous exposure to UV light, humidity, high-temperature changes, etc., can result in their degradation, unintended cross-linking and/or chain scission reactions that diminish protection, cause yellowing, or cause the polymeric layers to separate. To improve the coating qualities and durability, silicon-based and silane-based polymers were also evaluated with and without the addition of inorganic silica (SiO2) and titanium dioxide (TiO2) nanoparticles [9,90].
A hybrid coating, made of acrylate monomers (as binding agent) and chemical compounds such as organosilanes, fluorinated silanes and titania nanoparticles, has the necessary qualities, such as thermal resistance, mechanical resistance, weathering resistance, hydrophobicity and self-cleaning, to be used as a protective coating for cultural treasures made of precious stones. In this regard, methyl methacrylate and 3-(trimethoxysilyl)propyl methacrylate were employed for the preparation of a functional acrylate-based coating. In order to increase the coating’s thermal resistance, tetraethyl orthosilicate (TEOS, an organosilane) was added to the polymeric blend. Moreover, to achieve better hydrophobicity and resilience to weathering, perfluorooctyl-trichlorosilane is further added to the mixture. Additionally, the formulation was completed by adding titanium dioxide employed to enhance the coating’s thermal resistance and photocatalytic properties. The as-obtained multifunctional organic–inorganic hybrid nanocomposite was finally employed to coat a little area of historical monuments of Persepolis in Fars province (Iran) via impregnation. Cracking resistance, hydrophobicity, water absorption resistance, weathering durability, and color durability are only a few of the qualities that have been significantly improved by the presence of opportunely added functional compounds. The hydrophobicity of the coating is improved thanks to the achievement of a rough surface obtained by the employment of the organofluorine–titania hybrid nanocomposite. Additionally, titanium dioxide enhances thermal resistance and hardness and trigger photocatalytic activity, which results in the surface’s ability to self-clean [63].
To replace fossil-derived polymers and fluorine-compounds, some other more sustainable approaches and formulations can be evaluated. An example is represented by a silane/siloxane emulsion employed as a water repellent agent coupled with chitosan and silver nitrate as biocides to create an eco-friendly finishing with both hydrophobic and biocide qualities. Chitosan is a naturally occurring amino-polysaccharide obtained from the deacetylation reaction of chitin coming from the shells of crustaceans, and it is the second-most common biopolymer in nature after cellulose. It is distinguished by a high concentration of amine and hydroxyl functionalities and has attracted growing interest in several fields [128,129,130]. The formulation Tegosivin® HE 328 serves as the foundation for the protective coatings described in the example. It is an emulsion concentration built on alkoxy-functional silanes and organo-modified siloxanes. Chitosan was first added to the water repellent sol–gel formulation in various quantities as well as low concentrations of silver nitrate. A limestone called the Dom stone was finally spray-coated with the functional biocide and hydrophobic formulation achieving from 122.8° to 129.4° of WCA and a significant Chlorella vulgaris biocide impact [64].
Among the nanostructured hybrid coatings approach, some natural derivatives such as Zein can be employed to achieve hydrophobic protective coatings for cultural heritages. It is an amphiphilic prolamine obtained from corn endosperm and is characteristic of about 80% of corn proteins. Its hydrophilic behavior is determined by the relatively high content of glutamine (21–26%). A 5% (w/v) solution of zein in DMSO was therefore prepared to obtain a protective hydrophobic coating of a Serena stone (a fine-medium grain sandstone) with the spray coating technique. A WCA of about 120° was achieved. The purposed hydrophobic behavior mechanism of the obtained coating is shown in Figure 11. In particular, using spray coating, the solution was released as tiny droplets in the air, causing the solvent to quickly evaporate from the surface of the droplets. Before the solvent completely evaporated, a radial zein concentration gradient formed within each droplet in a very short amount of time. Zein began to harden, in particular, from the exterior of the droplet at the air–liquid interface toward the interior. The hydrophilic polar side of zein remains in contact with the solvent since there is still DMSO inside the droplet, forcing the non-polar hydrophobic side to face the outer portion of the droplets. As the drying process progresses, the solvent completely evaporates from the inside of the droplets as well. Atmospheric pressure placed on the partially solidified zein particles and the force of the droplets colliding with the stone surface cause the droplets to collapse on top of each other, creating the hydrophobic coating [65].

2.3.2. Functionalization Approaches for Textiles Surface Modification

To give or enhance functional qualities of common polymer fabrics, two main approaches could be employed. The first is centered on the creation of novel fibers, which is still an expensive strategy that frequently necessitates learning new techniques for producing products and acquiring new machinery for various materials. The second strategy is based on the more useful functionalization of the surface of traditional fibers or fabrics while making minimal changes to the manufacturing methods [131]. This latter strategy is the most intriguing in the field of functional and engineered textiles. In addition to functional hi-tech fabrics, the development of more sophisticated coatings has produced engineered textiles able to interact with and react to their surroundings for applications in different innovative and nanotechnological sectors (Figure 12).
The preparation of active coatings for smart and functional textiles can be achieved using a variety of methods, most of which are traditional, while others are more modern and cutting-edge, including nanotechnologies [132,133,134]. While all of these processes are based on the attachment of certain chemical moieties to textile surfaces, they differ in terms of the modifying substance, substrate to be changed, kind of modification, and other aspects. The fabric’s wettability and the coating’s adherence to the fabric sample are both important aspects of the coating process. While the adherence of the coating is necessary for stable contact with fabric surfaces, the wettability of the textiles by the coating has a specific impact on the wicking of the treated fabrics. Intermolecular forces, ionic or covalent bonds, or weak interactions (dipole–dipole, hydrogen bond, induced dipole–dipole interactions, or dispersive) may all play a role in coating adhesion to textiles [135].
The coating process optimization must be considered since it can affect a variety of important textile qualities, such as comfort, breathability, and the hand of the coated fabric. Furthermore, the surface of the fabric and the presence of impurities can both affect the coating adhesion [136]. To achieve both functional and smart textiles with outstanding performance, uniformity and good coating distribution on the fabric surface are crucial components.
After preparing the substrate, the next step is the designed synthesis of the coating, which can be represented by a combination of functional substances in solvents or emulsions. Stabilizing additives are frequently used in this step to keep the coating solution stable for the duration of the application process. A post-processing phase may be necessary after the coating application, such as curing (cross-linking) using thermal treatment or various energy sources, including gas, an infrared oven, or UV radiation [137,138].
The sol–gel reaction, fabric impregnation process, chemical grafting, layer-by-layer (LBL) assembly, and other surface modification technologies are examples [139,140,141]. Among the aforementioned fabric finishing techniques, sol–gel technology has been widely used and adopted in a variety of industries due to its gentle reaction conditions, high efficiency, cheap cost, and environmental friendliness [142]. As a matter of fact, the sol–gel technique represents an eco-friendly method of functionalizing fabric surfaces by depositing a thin layer with specific physical properties, chemical stability and optical transparency, thus representing an advantageous technology over the ones currently available for the introduction of specific functionality onto textile fabrics. In fact, the sol–gel method makes it feasible to create textile coatings with different implemented properties [57,143,144,145,146,147,148].
The widely used sol–gel coating application procedures are based on dip-coating, padding, or spraying, producing smart or functional textiles with precisely designed properties. The dip-pad-cure method was shown to be the most popular due to its simplicity and viability from an economic standpoint. Using a padder, the fabric is immersed or dipped into a coating material solution while moving at a constant speed. After drying and curing, the process is repeated (Figure 13).
The types of chemical bonds, involving the adhesion of a coating to a surface, include covalent bonds (such as those between a silane end and an OH-group belonging to a cotton cellulose molecule), molecular attractions (such as Van der Waals forces, hydrogen bonds, dipole–dipole interactions) and, finally, the retention of the molecule by the substrate through adhesive and cohesive forces between the molecule and the substrate, as well as the molecule to itself [103,149].
Recent years have seen the application of hydrophobic treatment to textile surfaces for antifouling, self-cleaning, anti-ice, and oil–water separation purposes [150,151]. Additionally, due to their low surface energy and oil/water repellent qualities, stain-resistant surfaces and anti-stain coatings have important applications in a variety of industries, including textiles, construction, cars, and electronics [152]. Unfortunately, fluoroalkyl silanes and fluorine compounds were frequently used to further increase the surface water repellency of textile fabrics [153,154]. The ECHA committees most recently proposed restricting the use of particular perfluoroalkyl compounds (PFAS) in certain application sectors [155]. As a result, examples of environmentally friendly, fluorine-free textile finishings with stain- and water-repellent properties are also documented [156,157,158].
For example, some TEOS and citric acid superhydrophobic coatings were developed using a sol–gel approach and a spray coating/dry-curing technique to achieve fabrics (90% cotton and 10% polyester) with a WCA above 150°. The strength of adhesion between the silica and the cotton was improved by the use of citric acid [66].
Moreover, functional alkyl(trialkoxy)silane-modified hybrid nanostructured materials were successfully produced and used as hydrophobic and water-based strain resistance coatings for cotton fabrics, via the sol–gel process and pad-dry-cure technique (Figure 14a,b).
The specific objective of the reported example was to investigate different functional alkyl(trialkoxy)silanes (Hexadecyltrimethoxysilane, Triethoxy(octyl)silane and Triethoxy(ethyl)silane: Cx and Cy) as precursors to synthesize efficient and stable hybrid sol–gel GPTMS-based coatings and further reduce cotton surface energy to enhance textile hydrophobicity and water-based stain resistance. By pad-dry-cure deposition of the produced nano-hybrid coatings, the double coating synthetic approach was successfully applied to increase the cotton surface’s hydrophobicity.
The hydrophobicity of the fabrics was evaluated by WCA measurements, which showed that the treated fabrics have high static contact angle values (up to roughly 150°). The resistance of the treated fabric to water-based stains against several tested liquids, solutions, and soil was also proved. The moisture adsorption analysis and the air permeability test were also used to assess the fabric quality, and the results of these tests revealed that coated cotton fabrics had better overall breathability compared to pristine cotton ones. According to all experimental results, the functional alkyl(trialkoxy)silane used in the sol–gel nanohybrid coatings provided treated cotton fabrics with excellent hydrophobicity and, as a result, water repellency through the synergistic action of the surfaces’ rough structure and low surface energy according to the mechanism purposed in Figure 15.
Among sol–gel processes and silane-compounds, other sustainable approaches and polymers have been recently employed for the production of hydrophobic and multifunctional coatings for textile materials.
Polyurethanes (PUs) are a type of versatile polymer that may easily have their structure and morphology adjusted to display a wide range of mechanical, physical, chemical, and biological characteristics. Foams, thin films, textiles, technology, automotives and buildings are just a few of the industrial uses for PUs [159]. In addition to offering advantages in terms of superior material properties (such as versatility, stretchability, elasticity, and mechanical strength), processability (such as 3D printing, inkjet printing, screen printing, spray coating, and molding), and sustainability, water-based polyurethanes are constantly developed in order to replace common solvent-based polyurethanes, but also other fossil-derived polymers [160].
An example is represented by a nanocomposite coating made of single-walled carbon nanotubes (SWCNT) and a series of waterborne polyurethane-urea dispersions (WPUD) free from tin catalysts and harmful solvents. The obtained WPUD are made of 100% bio-based semi-crystalline polyester polyol (Priplast 3294) and isophorone diisocyanate and are characterized by a bio-based content ranging from 59% to 72%. The functional WPUD coating with 0.1% of SWCNT was applied by knife-coating on polyester fabrics, obtaining a hydrophobic and conductive surface [68].
Citric acid (CTA), a bio-sourced acid, and polyethylene glycol (PEG200), a polyglycol, were combined in another example to create a sustainable functional polyol (SFP) (Figure 16a) employed for the production of a waterborne polyurethane coating (Figure 16b) for cotton fabrics in order to improve the antibacterial and breathable waterproof features. The waterborne polyurethane-coated samples obtained with a knife and the roller coating approach revealed inherent antibacterial capabilities ranging from 84% to 99% against Escherichia coli germs and water barrier qualities without a compromise in the mechanical properties of the cotton fabric [69].

2.4. Anti-Fouling and Fouling-Release Sustainable Coatings

The phenomenon of fouling occurs when macromolecules, bacteria or suspended particles stick to the surface of different materials. The construction industry, cultural heritage, marine industry and industrial sectors are all interested in this problem [161,162]. In this regard, one critical use of nanostructured coatings might result in the inhibition of microbial proliferation on various surfaces. Due to the variety of fouling organisms and their adhesion processes, designing antifouling (AF) coatings has proven to be a considerable challenge over time. Significant research on biocidal and non-biocidal coatings that prevent and decrease biofouling was sparked by the need to find answers to these problems [163,164]. Most control measures in the middle of the 19th century employed paints that included biocides. However, the biocide is a highly harmful agent (similar to the now-banned tributyltin) that is put into paints as an additive and, when exposed to seawater, is gradually released into the marine environment as a result of chemical and physical phenomena [165].
Fouling is influenced by surface characteristics, such as surface energy and wettability. As a result, altering the surface morphology and functionalities offers a key method for providing antifouling features on different surfaces [166]. One efficient way to do this is to treat the substrate with a formulation made of antifouling and antibacterial polymers. These coatings and paints can prevent biofouling and biocorrosion because they incorporate functional anti-adhesion, antimicrobial and anticorrosion chains rather than releasing biocides (Figure 17a–f).
In the next paragraphs, two main approaches of antifouling (AF) and fouling release (FR) (Figure 18) coating preparation will be explored, including the change in surface topography/hydrophilicity and the use of eco-friendly biocidal agents.

2.4.1. Biocide-Free FR Sustainable Coatings

All surfaces and equipment that operate in marine environments (such as small or large boats, pylons, mining platforms, and underwater monitoring systems) are regularly exposed to the corrosive action of environmental and biological agents, referred as marine fouling [168]. Sol–gel coating technology is probably one of the most pertinent tools for the development of environmentally friendly antifouling (AF) and fouling release (FR) formulations, given the current social expectation that new clean, flexible, and effective solutions will be adopted rather than the previously employed harmful ones. Due to several limitations of utilizing biocides in antifouling coatings, numerous techniques have been developed to obtain innovative coatings designed to directly interfere with microbe adherence as a result of topography or surface chemistry changes, thus showing FR properties.
As a matter of fact, by mixing silicones with fluoropolymers, the critical surface tension can be reduced. Additionally, extensive research has been conducted on developing hydrophobic surfaces that can prevent the onset of microfouling settlements [169]. Fluoropolymers can be successfully used to create anti-adhesion surfaces due to exposed CF2 and CF3 groups at the interface, which can limit the fouling attachment depending on the degree of fluorine atom mobility [170]. These fluorinated coatings provide good antibacterial action against a variety of microorganisms and negligible adhesive properties. Furthermore, tests on microbial cells showed that the substance discharged in the liquid medium had no biocidal effects. This non-biocidal approach is mostly suitable for vessels that travel at moderately high speeds and are not idle for extended periods of time, notwithstanding the high costs of fluoropolymer-modified finishings.
On this regard, four different silane precursors were used to combine the chemical-physical properties of co-monomers containing functional fluorinated organic compounds (i.e., 3,3,3-trifluoropropyl-trimethoxysilane, F3, and glycidyl-2,2,3,3,4,4,5,5,6,6,7,7,8,8,9,9-hexadecafluorononylether, F16) with those of alkoxysilane cross-linkers subunits containing epoxy and amine groups (i.e., GPTMS and APTES) to obtain a biocide-free FR and hydrophobic coating for the marine environment. Moreover, an asymmetric nanostructured hyperbranched polymeric film was produced by mixing both F3 and F16 co-monomers, i.e., bearing both long and short per-fluorinated chains [70].
For example, different painting/coating cycles are often employed in the naval sector surfaces in order to ensure adequate antifouling protection; they involve the use of three layers of polymeric AF/FR protective formulations on the hull surface exposed to the marine surrounding and are specifically indicated as primer, tie coat, and final topcoat. In another study, a research study was carried out on the design and development of functional hybrid coatings (top-coats) based on sol–gel approaches and featuring strong hydrophobic and antifouling/fouling-release capabilities. FR alternative systems were widely utilized in the biofouling mitigation treatment of various surfaces, offering a non-toxic substitute to AF coatings based on biocide- and fluorine-containing formulations. In more detail, the sol–gel cross-linker GPTMS precursor was used in conjunction with long alkyl-chain alkoxysilanes with various hydrocarbon chain lengths and hydro-repellent features (Figure 19).
This investigation showed that the surface wettability was significantly affected by these opportune organic silanes bearing suitable functional groups, such as the long alkyl chains. In addition, the coating’s hydrophobic properties were improved by the application of a commercial tie-coat layer between the primer and the topcoat, which decreased surface energy and raised the contact angle value. As a result, these eco-friendly coatings opened up the way to the development of fouling release sol–gel-based paints. By lowering the adhesion force between the fouler and the material used to construct the ship surface, their application could enable the fouling to be removed only by the movement of the ship during navigation [71].
A lotus leaf-inspired superhydrophobic coating was developed to fight microbiologically influenced corrosion in seawater environments. Copper surfaces were coated by a simple electrodeposition strategy based on a coordination-deposition process of a matrix made of an Fe-myristic acid compound from an ethanol solution in a single step. Therefore, a porous structure was developed starting from these hydrophobic compounds, which showed a dependent link between wettability and electrodeposition parameters such as electrodeposition potential and time. Due to the restricted anchoring sites and low surface energy, the as-obtained superhydrophobic matrix is verified to prevent representative diatoms and sulfate-reducing bacteria (SRB) from growing onto metal surfaces. In severe SRB suspension, the superhydrophobic matrix demonstrates remarkable corrosion resistance, suggesting a key role in combating microbiologically driven corrosion in saltwater environments [72].
AF/FR coatings can also be developed using natural precursors such as curcumin. In particular, a bio-based antifouling and anticorrosion benzoxazine double-layer coating for metal surfaces was obtained from the reaction of curcumin, APTES and paraformaldehyde (Figure 20a). The synthesized benzoxazine was loaded with various amounts of hydrophilic PEG molecules to modify the resin framework, give the resin its intrinsic characteristics and improve anti-biofouling performance. Curcumin has several inherent properties; moreover, the two phenolic hydroxyl groups characterizing the molecule can increase the cross-link density of the bioactive bisbenzoxazine resin (Figure 20b). Furthermore, APTES is utilized as a common and inexpensive amine source to enhance the resin cross-linking (Si-O-Si framework formation). The resins exhibited good repellent properties (prevention of settlement, adhesion, and growth of microorganisms) even in the absence of booster biocides, thanks to the advantageous effects of specific PEG content [73].
In another work, different bio-based polybenzoxazine/copolymer coatings, containing arbutin and silane functionalization featuring anticorrosive and anti-biofouling properties for low-carbon steel substrates, were developed. In detail, the plant-derived arbutin (β-d-glucoside of hydroquinone) was employed in synergic combination with the silane precursor [3-(2-aminoethylamino) propyl]trimethoxysilane (AEAPTMS) and paraformaldehyde for the production of the benzoxazine polymer resin (Figure 21a,b) applied by dip-coating on the selected substrates. Arbutin and its many hydroxyl groups contribute to its ability to enhance cross-linking. PEG or PEG-based copolymers are also added to increase surface resistance to antimicrobial adhesion. In this regard, polydopamine-PEG coating seemed to work as both an anti-bio adhesive and an anti-biofouling agent, reducing biofouling attachment on a variety of surfaces [74].

2.4.2. Eco-Friendly Biocide AF Coatings

Biocide-containing AF coatings are also developed by following sustainable procedures and approaches or by employing nonharmful substances for the environment and human health. In this regard, inorganic nanoparticles can find applications as biocidal agents for antifouling applications. Since the restriction on organotin chemicals, copper has become the most significant alternative biocide for antifouling applications [165]. The methods used to incorporate copper into polymeric matrices need to be further investigated to ensure the controlled release of copper ions without compromising the antifouling performances. The development of copper-containing polymer coatings using the liquid flame spray technique is an example. This technique allows the one-step preparation of polymer-encapsulated composite coatings, starting from proper formulations and dispersions containing different additives. For this purpose, copper nanoparticles were incorporated into a polyimide precursor to produce micro-structured polyimide-copper splats and coatings for the controlled release of copper and the inhibition of the tested Bacillus sp., Phaeodactylum tricornutum and Chlorella foulants. In particular, a Kapton-type aromatic polyimide precursor solution was obtained by mixing the monomer pyromellitic dianhydride (PMDA) and 4,4′-oxydianiline (ODA) (Figure 22a), and subsequently, the polyimide-Cu suspension was prepared by adding Cu nanoparticles (~300 nm in size). Carbon steel Q235 plates were finally coated with the liquid flame spray approach with the prepared formulation (Figure 22b) [75].
Other approaches and polymeric formulations can also be developed for the immobilization of antifouling agents such as the isocyanate-reactive Econea biocide (4-bromo-2-(4-chlorophenyl)-5-(trifluoromethyl)-1H-pyrrole-3-carbonitrile), created by functionalizing econea with a diisocyanate, to avoid their dispersion into the environment. When compared to other typical biocide-releasing coatings, this latter biocide was successfully grafted onto polydimethylsiloxane (PDMS)-based coatings, thus resulting in a tenfold decrease in the leaching qualities of the created coatings into the aquatic environment.
The multi-resistant Staphylococcus aureus strain was successfully inhibited by the PDMS-based coatings containing grafted Econea, which also exhibited favorable antibacterial properties and bacteriostatic activity. The coating developed on acrylic prototypes remained cleaner than the free biocide reference antifouling coating after both were submerged in Atlantic seawater for a duration of 30 months [76].
Moreover, some nanocarriers, such as halloysite clay nanotubes, can be employed for the encapsulation and controlled slow release of biocides. Because of its tubular shape, halloysite can also be loaded with various compounds to be employed for the release of active molecules, such as antioxidants, flame retardants, corrosion inhibitors, biocides, and medicines [171,172,173]. For these purposes, two eco-sustainable antifouling compounds, sodium salicylate and N-(2,4,6-trichlorophenyl)maleimide, were successfully encapsulated into HNTs and used for the development of a hybrid composite based on epoxy paint (Figure 23a,b).
The electrostatic interaction between the cavities of HNTs and the chosen active molecules and the solvent for their dissolution was properly tuned in order to achieve higher loading values. After the etching of HNTs, the nanotubes were loaded with the two different fillers (Figure 23a) and incorporated (10 wt% of functionalized HNT) into a commercial epoxy paint and the relative amide-based curing agent. The resulting hybrid paint was therefore poured on smooth plastic sheets with a film-coating approach. These experiments revealed that the modest and constant release of the active molecules was sufficient to stop the adhesion of Vibrio natriegens marine bacterium on the coated substrates (Figure 23b) [77].

3. Final Remarks and Future Perspectives

The growing need for more performant and long-life materials is going to have to deal with the impact on the environment of the actual technologies and approaches for the production of protective coatings and finishings. In this regard, new and more sustainable methods are constantly developed in order to reduce or replace the common fossil-derived polymers, waxes and potentially harmful compounds actually still employed for the improvement and preservation of surfaces from corrosion, fire, bacteria, water and other external and environmental degrading agents.
In this context, this review aims to give an outlook on the most recent methodologies for the eco-friendly design and development of metal, plastic, textile and stone surfaces for protective coatings and finishings. The main applications of the latter rely on the implementation of the material features, among which is their anticorrosive, flame-retardant, hydrophobic, antibacterial, antifouling and fouling-release behavior. Different synthetic protocols and approaches, such as sol–gel synthesis, are evaluated and compared. Additionally, the use of naturally derived compounds, waste materials and waterborne formulations for sustainable coatings preparation is described. Different innovative application techniques are shown to achieve a better homogeneity of the coating and nano/micro-structured textures and morphologies.
Sol–gel-based matrices, organic–inorganic hybrid materials, waterborne polyurethanes, bio-based benzoxazine, and nanocarriers for active substances are among the examples shown for achieving a solution to these challenges and thus attempting the need for less-impacting and safe coating formulations. The scientific advancements in the area of sustainable nanostructured coatings highlighted in this study show the significance of the role that nanotechnology can play in the production of increasingly high-value and competitive coatings and finishings for different application sectors. Table 2 reports a comparison between each strategy described in this review, mainly focusing on their advantages and disadvantages.
However, the development of even more advanced polymers and formulations that can provide surfaces with permanent or long-lasting multifunctional qualities still poses some difficulties. For instance, a multi-layer approach could be employed for surface preparation and functionalization. The advantages of this approach rely on a better adhesion of the final functional coating to the pre-treated surface (i.e., with proper tie-coats and primer employment) but also on a higher exposure to the active sites of the specimen’s functional coating. Other problems come from the attempts to reduce the environmental impact of the finishing procedures and to implement these eco-friendly coatings on an industrial scale with the aim to be commercialized in the reference market when applied to several materials and surfaces of everyday life. In this context, a multidisciplinary strategy is required to attain these goals. As a result, collaboration across many scientific areas and research activities involved in sustainable coating manufacturing for various applications may be a solution to these challenges, thus resulting in more performant and improved surfaces.

Author Contributions

Conceptualization, G.R., S.S. and M.R.P.; resources, M.R.P.; data curation, G.R., S.S. and M.R.P.; writing—original draft preparation, G.R., S.S. and M.R.P.; writing—review and editing, G.R., S.S. and M.R.P.; supervision, M.R.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

MURST: CNR and MUR are gratefully acknowledged for financial support. All authors wish to thank S. Romeo, G. Napoli and F. Giordano for technical and informatics assistance. This work was carried out as part of G.R.’s PhD program, supported by PON-MUR “Ricerca e Innovazione 2014–2020” RESTART project.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhang, H.; Akhtar, F. Refractory Multicomponent Boron-Carbide High Entropy Oxidation-Protective Coating for Carbon-Carbon Composites. Surf. Coat. Technol. 2021, 425, 127697. [Google Scholar] [CrossRef]
  2. Awang, M.; Khalili, A.A.; Pedapati, S.R. A Review: Thin Protective Coating for Wear Protection in High-Temperature Application. Metals 2020, 10, 42. [Google Scholar] [CrossRef] [Green Version]
  3. Zhu, Q.; Chua, M.H.; Ong, P.J.; Cheng Lee, J.J.; Le Osmund Chin, K.; Wang, S.; Kai, D.; Ji, R.; Kong, J.; Dong, Z.; et al. Recent Advances in Nanotechnology-Based Functional Coatings for the Built Environment. Mater. Today Adv. 2022, 15, 100270. [Google Scholar] [CrossRef]
  4. Tejero-Martin, D.; Rezvani Rad, M.; McDonald, A.; Hussain, T. Beyond Traditional Coatings: A Review on Thermal-Sprayed Functional and Smart Coatings. J. Therm. Spray Technol. 2019, 28, 598–644. [Google Scholar] [CrossRef] [Green Version]
  5. Dhanyalayam, D.; Scrivano, L.; Parisi, O.I.; Sinicropi, M.S.; Fazio, A.; Saturnino, C.; Plutino, M.R.; Di Cristo, F.; Puoci, F.; Cappello, A.R.; et al. Biopolymeric Self-Assembled Nanoparticles for Enhanced Antibacterial Activity of Ag-Based Compounds. Int. J. Pharm. 2017, 517, 395–402. [Google Scholar] [CrossRef] [PubMed]
  6. Rebelo, P.; Costa-Rama, E.; Seguro, I.; Pacheco, J.G.; Nouws, H.P.A.; Cordeiro, M.N.D.S.; Delerue-Matos, C. Molecularly Imprinted Polymer-Based Electrochemical Sensors for Environmental Analysis. Biosens. Bioelectron. 2021, 172, 112719. [Google Scholar] [CrossRef]
  7. Trovato, V.; Sfameni, S.; Rando, G.; Rosace, G.; Libertino, S.; Ferri, A.; Plutino, M.R. A Review of Stimuli-Responsive Smart Materials for Wearable Technology in Healthcare: Retrospective, Perspective, and Prospective. Molecules 2022, 27, 5709. [Google Scholar] [CrossRef] [PubMed]
  8. Thiel, C.; Neumann, K.; Ludwar, F.; Rennings, A.; Doose, J.; Erni, D. Coating Damage Localization of Naval Vessels Using Artificial Neural Networks. Ocean Eng. 2019, 192, 106560. [Google Scholar] [CrossRef]
  9. Ielo, I.; Galletta, M.; Rando, G.; Sfameni, S.; Cardiano, P.; Sabatino, G.; Drommi, D.; Rosace, G.; Plutino, M.R. Design, Synthesis and Characterization of Hybrid Coatings Suitable for Geopolymeric-Based Supports for the Restoration of Cultural Heritage. IOP Conf. Ser. Mater. Sci. Eng. 2020, 777, 012003. [Google Scholar] [CrossRef]
  10. Krzak, J.; Szczurek, A.; Babiarczuk, B.; Gąsiorek, J.; Borak, B. Chapter 5—Sol–Gel Surface Functionalization Regardless of Form and Type of Substrate. In Micro and Nano Technologies; Elsevier: Amsterdam, The Netherlands, 2020; pp. 111–147. ISBN 978-0-12-821381-0. [Google Scholar]
  11. De Pasquale, G.; Sibona, S. Hybrid Materials Based on Polymers-Filled AM Steel Lattices with Energy Absorption Capabilities. Mech. Adv. Mater. Struct. 2022, 29, 2570–2580. [Google Scholar] [CrossRef]
  12. Tran, M.H.; Brilmayer, R.; Liu, L.; Zhuang, H.; Hess, C.; Andrieu-Brunsen, A.; Birkel, C.S. Synthesis of a Smart Hybrid MXene with Switchable Conductivity for Temperature Sensing. ACS Appl. Nano Mater. 2020, 3, 4069–4076. [Google Scholar] [CrossRef]
  13. Liras, M.; Barawi, M.; de la Peña O’Shea, V.A. Hybrid Materials Based on Conjugated Polymers and Inorganic Semiconductors as Photocatalysts: From Environmental to Energy Applications. Chem. Soc. Rev. 2019, 48, 5454–5487. [Google Scholar] [CrossRef] [PubMed]
  14. Ashraf, M.-T.; AlHammadi, A.A.; El-Sherbeeny, A.M.; Alhammadi, S.; Al Zoubi, W.; Ko, Y.G.; Abukhadra, M.R. Synthesis of Cellulose Fibers/Zeolite-A Nanocomposite as an Environmental Adsorbent for Organic and Inorganic Selenium Ions; Characterization and Advanced Equilibrium Studies. J. Mol. Liq. 2022, 360, 119573. [Google Scholar] [CrossRef]
  15. Rando, G.; Sfameni, S.; Galletta, M.; Drommi, D.; Cappello, S.; Plutino, M.R. Functional Nanohybrids and Nanocomposites Development for the Removal of Environmental Pollutants and Bioremediation. Molecules 2022, 27, 4856. [Google Scholar] [CrossRef]
  16. Ielo, I.; Iacopetta, D.; Saturnino, C.; Longo, P.; Galletta, M.; Drommi, D.; Rosace, G.; Sinicropi, M.S.; Plutino, M.R. Gold Derivatives Development as Prospective Anticancer Drugs for Breast Cancer Treatment. Appl. Sci. 2021, 11, 2089. [Google Scholar] [CrossRef]
  17. Saturnino, C.; Caruso, A.; Longo, P.; Capasso, A.; Pingitore, A.; Caroleo, M.C.; Cione, E.; Perri, M.; Nicolo, F.; Nardo, V.M.; et al. Crystallographic Study and Biological Evaluation of 1,4-Dimethyl-N-Alkylcarbazoles&#8224. Curr. Top. Med. Chem. 2015, 15, 973–979. [Google Scholar]
  18. Iacopetta, D.; Catalano, A.; Ceramella, J.; Saturnino, C.; Salvagno, L.; Ielo, I.; Drommi, D.; Scali, E.; Plutino, M.R.; Rosace, G.; et al. The Different Facets of Triclocarban: A Review. Molecules 2021, 26, 2811. [Google Scholar] [CrossRef]
  19. Artesani, A.; Di Turo, F.; Zucchelli, M.; Traviglia, A. Recent Advances in Protective Coatings for Cultural Heritage–An Overview. Coatings 2020, 10, 217. [Google Scholar] [CrossRef] [Green Version]
  20. Chen, S.; Li, L.; Zhao, C.; Zheng, J. Surface Hydration: Principles and Applications toward Low-Fouling/Nonfouling Biomaterials. Polymer 2010, 51, 5283–5293. [Google Scholar] [CrossRef] [Green Version]
  21. Liu, Y.; Wang, X.; Qi, K.; Xin, J.H. Functionalization of Cotton with Carbon Nanotubes. J. Mater. Chem. 2008, 18, 3454. [Google Scholar] [CrossRef]
  22. Catauro, M.; Vecchio Ciprioti, S. Sol-Gel Synthesis and Characterization of Hybrid Materials for Biomedical Applications BT. In Thermodynamics and Biophysics of Biomedical Nanosystems: Applications and Practical Considerations; Demetzos, C., Pippa, N., Eds.; Springer: Singapore, 2019; pp. 445–475. ISBN 978-981-13-0989-2. [Google Scholar]
  23. Saturnino, C.; Popolo, A.; Ramunno, A.; Adesso, S.; Pecoraro, M.; Plutino, M.R.; Rizzato, S.; Albinati, A.; Marzocco, S.; Sala, M.; et al. Anti-Inflammatory, Antioxidant and Crystallographic Studies of N-Palmitoyl-Ethanol Amine (PEA) Derivatives. Molecules 2017, 22, 616. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Wei, H.; Wu, F.; Li, L.; Yang, X.; Xu, C.; Yu, P.; Ma, F.; Mao, L. Natural Leukocyte Membrane-Masked Microelectrodes with an Enhanced Antifouling Ability and Biocompatibility for In Vivo Electrochemical Sensing. Anal. Chem. 2020, 92, 11374–11379. [Google Scholar] [CrossRef]
  25. Sun, W.; Wu, T.; Wang, L.; Yang, Z.; Zhu, T.; Dong, C.; Liu, G. The Role of Graphene Loading on the Corrosion-Promotion Activity of Graphene/Epoxy Nanocomposite Coatings. Compos. Part B Eng. 2019, 173, 106916. [Google Scholar] [CrossRef]
  26. Mansfield, E.; Tyner, K.M.; Poling, C.M.; Blacklock, J.L. Determination of Nanoparticle Surface Coatings and Nanoparticle Purity Using Microscale Thermogravimetric Analysis. Anal. Chem. 2014, 86, 1478–1484. [Google Scholar] [CrossRef]
  27. Zheng, S.; Chen, H.; Tong, X.; Wang, Z.; Crittenden, J.C.; Huang, M. Integration of a Photo-Fenton Reaction and a Membrane Filtration Using CS/PAN@FeOOH/g-C3N4Electrospun Nanofibers: Synthesis, Characterization, Self-Cleaning Performance and Mechanism. Appl. Catal. B Environ. 2021, 281, 119519. [Google Scholar] [CrossRef]
  28. Ilyas, R.A.; Sapuan, S.M.; Bayraktar, E. Current Progress in Biopolymer-Based Bionanocomposites and Hybrid Materials. Polymers 2022, 14, 3479. [Google Scholar] [CrossRef]
  29. Lei, Q.; Guo, J.; Noureddine, A.; Wang, A.; Wuttke, S.; Brinker, C.J.; Zhu, W. Sol–Gel-Based Advanced Porous Silica Materials for Biomedical Applications. Adv. Funct. Mater. 2020, 30, 1909539. [Google Scholar] [CrossRef]
  30. Baino, F.; Fiume, E.; Miola, M.; Verné, E. Bioactive Sol-Gel Glasses: Processing, Properties, and Applications. Int. J. Appl. Ceram. Technol. 2018, 15, 841–860. [Google Scholar] [CrossRef]
  31. Sfameni, S.; Hadhri, M.; Rando, G.; Drommi, D.; Rosace, G.; Trovato, V.; Plutino, M.R. Inorganic Finishing for Textile Fabrics: Recent Advances in Wear-Resistant, UV Protection and Antimicrobial Treatments. Inorganics 2023, 11, 19. [Google Scholar] [CrossRef]
  32. Ben Chobba, M.; Weththimuni, M.L.; Messaoud, M.; Urzi, C.; Bouaziz, J.; De Leo, F.; Licchelli, M. Ag-TiO2/PDMS Nanocomposite Protective Coatings: Synthesis, Characterization, and Use as a Self-Cleaning and Antimicrobial Agent. Prog. Org. Coat. 2021, 158, 106342. [Google Scholar] [CrossRef]
  33. Huang, G.; Huo, L.; Jin, Y.; Yuan, S.; Zhao, R.; Zhao, J.; Li, Z.; Li, Y. Fluorine-Free Superhydrophobic PET Fabric with High Oil Flux for Oil–Water Separation. Prog. Org. Coat. 2022, 163, 106671. [Google Scholar] [CrossRef]
  34. Liu, C.; Xing, T.; Wei, B.; Chen, G. Synergistic Effects and Mechanism of Modified Silica Sol Flame Retardant Systems on Silk Fabric. Materials 2018, 11, 1842. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Rao, K.S.; El-Hami, K.; Kodaki, T.; Matsushige, K.; Makino, K. A Novel Method for Synthesis of Silica Nanoparticles. J. Colloid Interface Sci. 2005, 289, 125–131. [Google Scholar] [CrossRef] [PubMed]
  36. Pierre, A.C. Introduction to Sol-Gel Processing; Springer Nature: Berlin, Germany, 2020; ISBN 3030381447. [Google Scholar]
  37. Alexandru, M.; Cazacu, M.; Cristea, M.; Nistor, A.; Grigoras, C.; Simionescu, B.C. Poly(Siloxane-Urethane) Crosslinked Structures Obtained by Sol-Gel Technique. J. Polym. Sci. Part A Polym. Chem. 2011, 49, 1708–1718. [Google Scholar] [CrossRef]
  38. Elanany, M.; Selvam, P.; Yokosuka, T.; Takami, S.; Kubo, M.; Imamura, A.; Miyamoto, A. A Quantum Molecular Dynamics Simulation Study of the Initial Hydrolysis Step in Sol−Gel Process. J. Phys. Chem. B 2003, 107, 1518–1524. [Google Scholar] [CrossRef]
  39. Esposito, S. “Traditional” Sol-Gel Chemistry as a Powerful Tool for the Preparation of Supported Metal and Metal Oxide Catalysts. Materials 2019, 12, 668. [Google Scholar] [CrossRef] [Green Version]
  40. Nisticò, R.; Scalarone, D.; Magnacca, G. Sol-Gel Chemistry, Templating and Spin-Coating Deposition: A Combined Approach to Control in a Simple Way the Porosity of Inorganic Thin Films/Coatings. Microporous Mesoporous Mater. 2017, 248, 18–29. [Google Scholar] [CrossRef]
  41. Brinker, C.J.; Frye, G.C.; Hurd, A.J.; Ashley, C.S. Fundamentals of Sol-Gel Dip Coating. Thin Solid Films 1991, 201, 97–108. [Google Scholar] [CrossRef]
  42. Firdaus, C.M.; Rizam, M.S.B.S.; Rusop, M.; Hidayah, S.R. Characterization of ZnO and ZnO: TiO2 Thin Films Prepared by Sol-Gel Spray-Spin Coating Technique. Procedia Eng. 2012, 41, 1367–1373. [Google Scholar] [CrossRef]
  43. Butt, M.A. Thin-Film Coating Methods: A Successful Marriage of High-Quality and Cost-Effectiveness—A Brief Exploration. Coatings 2022, 12, 1115. [Google Scholar] [CrossRef]
  44. Wu, Q.; Qin, K.-X.; Gan, M.-X.; Xu, J.; Li, Z.-L.; Li, Z.-C. Recyclable Biomass-Derived Polyethylene-Like Materials as Functional Coatings for Commercial Fabrics: Toward Upcycling of Waste Textiles. ACS Sustain. Chem. Eng. 2022, 10, 17187–17197. [Google Scholar] [CrossRef]
  45. Vincenzo Omelan, M.C.; Lachmann, K.; Meyer, H.; Stieghan, A.-K.; Thomas, M. Bio-Based Thin Film Coatings Using Sustainable Materials. Procedia CIRP 2022, 110, 20–25. [Google Scholar] [CrossRef]
  46. Wang, J.; Qian, W.; He, Y.; Xiong, Y.; Song, P.; Wang, R.-M. Reutilization of Discarded Biomass for Preparing Functional Polymer Materials. Waste Manag. 2017, 65, 11–21. [Google Scholar] [CrossRef] [PubMed]
  47. Shamaei, L.; Khorshidi, B.; Islam, M.A.; Sadrzadeh, M. Industrial Waste Lignin as an Antifouling Coating for the Treatment of Oily Wastewater: Creating Wealth from Waste. J. Clean. Prod. 2020, 256, 120304. [Google Scholar] [CrossRef]
  48. Hincapié, I.; Künniger, T.; Hischier, R.; Cervellati, D.; Nowack, B.; Som, C. Nanoparticles in Facade Coatings: A Survey of Industrial Experts on Functional and Environmental Benefits and Challenges. J. Nanopart. Res. 2015, 17, 287. [Google Scholar] [CrossRef]
  49. Cunningham, M.F.; Campbell, J.D.; Fu, Z.; Bohling, J.; Leroux, J.G.; Mabee, W.; Robert, T. Future Green Chemistry and Sustainability Needs in Polymeric Coatings. Green Chem. 2019, 21, 4919–4926. [Google Scholar] [CrossRef]
  50. Wypych, G. Handbook of Material Weathering; Elsevier: Amsterdam, The Netherlands, 2018; ISBN 1927885329. [Google Scholar]
  51. Riahinezhad, M.; Hallman, M.; Masson, J.-F. Critical Review of Polymeric Building Envelope Materials: Degradation, Durability and Service Life Prediction. Buildings 2021, 11, 299. [Google Scholar] [CrossRef]
  52. McMullan, R. Environmental Science in Building; Bloomsbury Publishing: London, UK, 2017; ISBN 1137605456. [Google Scholar]
  53. Nguyen-Tri, P.; Tran, H.N.; Plamondon, C.O.; Tuduri, L.; Vo, D.-V.N.; Nanda, S.; Mishra, A.; Chao, H.-P.; Bajpai, A.K. Recent Progress in the Preparation, Properties and Applications of Superhydrophobic Nano-Based Coatings and Surfaces: A Review. Prog. Org. Coat. 2019, 132, 235–256. [Google Scholar] [CrossRef]
  54. Riaz, S.; Ashraf, M.; Hussain, T.; Hussain, M.T.; Rehman, A.; Javid, A.; Iqbal, K.; Basit, A.; Aziz, H. Functional Finishing and Coloration of Textiles with Nanomaterials. Color. Technol. 2018, 134, 327–346. [Google Scholar] [CrossRef]
  55. Rabajczyk, A.; Zielecka, M.; Małozięć, D. Hazards Resulting from the Burning Wood Impregnated with Selected Chemical Compounds. Appl. Sci. 2020, 10, 6093. [Google Scholar] [CrossRef]
  56. Shammas, N.K.; Wang, L.K. Sulfide Precipitation for Treatment of Metal Wastes. In Handbook of Advanced Industrial and Hazardous Wastes Management; CRC Press: Boca Raton, FL, USA, 2017; pp. 185–238. ISBN 1315117428. [Google Scholar]
  57. Sfameni, S.; Del Tedesco, A.; Rando, G.; Truant, F.; Visco, A.; Plutino, M.R. Waterborne Eco-Sustainable Sol–Gel Coatings Based on Phytic Acid Intercalated Graphene Oxide for Corrosion Protection of Metallic Surfaces. Int. J. Mol. Sci. 2022, 23, 12021. [Google Scholar] [PubMed]
  58. Liu, S.; Liu, L.; Guo, H.; Oguzie, E.E.; Li, Y.; Wang, F. Electrochemical Polymerization of Polyaniline-Reduced Graphene Oxide Composite Coating on 5083 Al Alloy: Role of Reduced Graphene Oxide. Electrochem. Commun. 2019, 98, 110–114. [Google Scholar] [CrossRef]
  59. Jaramillo, A.F.; Montoya, L.F.; Prabhakar, J.M.; Sanhueza, J.P.; Fernández, K.; Rohwerder, M.; Rojas, D.; Montalba, C.; Melendrez, M.F. Formulation of a Multifunctional Coating Based on Polyphenols Extracted from the Pine Radiata Bark and Functionalized Zinc Oxide Nanoparticles: Evaluation of Hydrophobic and Anticorrosive Properties. Prog. Org. Coat. 2019, 135, 191–204. [Google Scholar] [CrossRef]
  60. Karattu Veedu, K.; Peringattu Kalarikkal, T.; Jayakumar, N.; Gopalan, N.K. Anticorrosive Performance of Mangifera Indica L. Leaf Extract-Based Hybrid Coating on Steel. ACS Omega 2019, 4, 10176–10184. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  61. Vahabi, H.; Jouyandeh, M.; Parpaite, T.; Saeb, M.R.; Ramakrishna, S. Coffee Wastes as Sustainable Flame Retardants for Polymer Materials. Coatings 2021, 11, 1021. [Google Scholar] [CrossRef]
  62. Attia, N.F.; Mohamed, A.; Hussein, A.; El-Demerdash, A.-G.M.; Kandil, S.H. Bio-Inspired One-Dimensional Based Textile Fabric Coating for Integrating High Flame Retardancy, Antibacterial, Toxic Gases Suppression, Antiviral and Reinforcement Properties. Polym. Degrad. Stab. 2022, 205, 110152. [Google Scholar] [CrossRef]
  63. Azadi, N.; Parsimehr, H.; Ershad-Langroudi, A. Cultural Heritage Protection via Hybrid Nanocomposite Coating. Plast. Rubber Compos. 2020, 49, 414–424. [Google Scholar] [CrossRef]
  64. Eyssautier-Chuine, S.; Calandra, I.; Vaillant-Gaveau, N.; Fronteau, G.; Thomachot-Schneider, C.; Hubert, J.; Pleck, J.; Gommeaux, M. A New Preventive Coating for Building Stones Mixing a Water Repellent and an Eco-Friendly Biocide. Prog. Org. Coat. 2018, 120, 132–142. [Google Scholar] [CrossRef]
  65. Zucchelli, M.; Mazzon, G.; Bertolacci, L.; Carzino, R.; Zendri, E.; Athanassiou, A. Stone Sustainable Protection and Preservation Using a Zein-Based Hydrophobic Coating. Prog. Org. Coat. 2021, 159, 106434. [Google Scholar] [CrossRef]
  66. Espanhol-Soares, M.; Costa, L.; Silva, M.R.A.; Soares Silva, F.; Ribeiro, L.M.S.; Gimenes, R. Super-Hydrophobic Coatings on Cotton Fabrics Using Sol–Gel Technique by Spray. J. Sol-Gel Sci. Technol. 2020, 95, 22–33. [Google Scholar] [CrossRef]
  67. Sfameni, S.; Lawnick, T.; Rando, G.; Visco, A.; Textor, T.; Plutino, M.R. Functional Silane-Based Nanohybrid Materials for the Development of Hydrophobic and Water-Based Stain Resistant Cotton Fabrics Coatings. Nanomaterials 2022, 12, 3404. [Google Scholar] [CrossRef] [PubMed]
  68. Lacruz, A.; Salvador, M.; Blanco, M.; Vidal, K.; Goitandia, A.M.; Martinková, L.; Kyselka, M.; de Ilarduya, A.M. Biobased Waterborne Polyurethane-Urea/SWCNT Nanocomposites for Hydrophobic and Electrically Conductive Textile Coatings. Polymers 2021, 13, 1624. [Google Scholar] [CrossRef]
  69. Bramhecha, I.; Sheikh, J. Development of Sustainable Citric Acid-Based Polyol To Synthesize Waterborne Polyurethane for Antibacterial and Breathable Waterproof Coating of Cotton Fabric. Ind. Eng. Chem. Res. 2019, 58, 21252–21261. [Google Scholar] [CrossRef]
  70. Sfameni, S.; Rando, G.; Galletta, M.; Ielo, I.; Brucale, M.; De Leo, F.; Cardiano, P.; Cappello, S.; Visco, A.; Trovato, V.; et al. Design and Development of Fluorinated and Biocide-Free Sol–Gel Based Hybrid Functional Coatings for Anti-Biofouling/Foul-Release Activity. Gels 2022, 8, 538. [Google Scholar] [CrossRef] [PubMed]
  71. Sfameni, S.; Rando, G.; Marchetta, A.; Scolaro, C.; Cappello, S.; Urzì, C.; Visco, A.; Plutino, M.R. Development of Eco-Friendly Hydrophobic and Fouling-Release Coatings for Blue-Growth Environmental Applications: Synthesis, Mechanical Characterization and Biological Activity. Gels 2022, 8, 528. [Google Scholar] [CrossRef] [PubMed]
  72. Li, B.; Ouyang, Y.; Haider, Z.; Zhu, Y.; Qiu, R.; Hu, S.; Niu, H.; Zhang, Y.; Chen, M. One-Step Electrochemical Deposition Leading to Superhydrophobic Matrix for Inhibiting Abiotic and Microbiologically Influenced Corrosion of Cu in Seawater Environment. Colloids Surf. A Physicochem. Eng. Asp. 2021, 616, 126337. [Google Scholar] [CrossRef]
  73. Deng, Y.; Xia, L.; Song, G.-L.; Zhao, Y.; Zhang, Y.; Xu, Y.; Zheng, D. Development of a Curcumin-Based Antifouling and Anticorrosion Sustainable Polybenzoxazine Resin Composite Coating. Compos. Part B Eng. 2021, 225, 109263. [Google Scholar] [CrossRef]
  74. Periyasamy, T.; Raorane, C.J.; Haldhar, R.; Asrafali, S.P.; Kim, S.-C. Development of Arbutin Based Sustainable Polybenzoxazine Resin for Antifouling and Anticorrosion of Low Carbon Steel. Prog. Org. Coat. 2022, 170, 106968. [Google Scholar] [CrossRef]
  75. Liu, Y.; Suo, X.; Wang, Z.; Gong, Y.; Wang, X.; Li, H. Developing Polyimide-Copper Antifouling Coatings with Capsule Structures for Sustainable Release of Copper. Mater. Des. 2017, 130, 285–293. [Google Scholar] [CrossRef]
  76. Ferreira, O.; Rijo, P.; Gomes, J.F.; Santos, R.; Monteiro, S.; Vilas-Boas, C.; Correia-da-Silva, M.; Almada, S.; Alves, L.G.; Bordado, J.C.; et al. Biofouling Inhibition with Grafted Econea Biocide: Toward a Nonreleasing Eco-Friendly Multiresistant Antifouling Coating. ACS Sustain. Chem. Eng. 2020, 8, 12–17. [Google Scholar] [CrossRef]
  77. Tonelli, M.; Perini, I.; Ridi, F.; Baglioni, P. Improving the Properties of Antifouling Hybrid Composites: The Use of Halloysites as Nano-Containers in Epoxy Coatings. Colloids Surf. A Physicochem. Eng. Asp. 2021, 623, 126779. [Google Scholar] [CrossRef]
  78. Marcus, P. Corrosion Mechanisms in Theory and Practice; CRC Press: Boca Raton, FL, USA, 2011; ISBN 1420094629. [Google Scholar]
  79. Gartner, N.; Kosec, T.; Legat, A. Monitoring the Corrosion of Steel in Concrete Exposed to a Marine Environment. Materials 2020, 13, 407. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Stansbury, E.E.; Buchanan, R.A. Fundamentals of Electrochemical Corrosion; ASM International: Novelty, OH, USA, 2000; ISBN 1615030670. [Google Scholar]
  81. Ianelli, R.F.; Saliba-Silva, A.M.; Takara, E.M.; Neto, J.G.; Souza, J.A.B.; Urano de Carvalho, E.F.; Durazzo, M. Nickel Electrodeposition in LEU Metal Foil Annular Targets to Produce Mo-99. Mater. Chem. Phys. 2022, 290, 126620. [Google Scholar] [CrossRef]
  82. Meng, Y.; Deng, J.; Ge, D.; Wu, J.; Sun, W.; Wang, R. Surface Textures Fabricated by Laser and Ultrasonic Rolling for Improving Tribological Properties of TiAlSiN Coatings. Tribol. Int. 2021, 164, 107248. [Google Scholar] [CrossRef]
  83. Pehkonen, S.O.; Yuan, S. Tailored Thin Coatings for Corrosion Inhibition Using a Molecular Approach; Academic Press: Cambridge, MA, USA, 2018; ISBN 0128135859. [Google Scholar]
  84. Ong, G.; Kasi, R.; Subramaniam, R. A Review on Plant Extracts as Natural Additives in Coating Applications. Prog. Org. Coat. 2021, 151, 106091. [Google Scholar] [CrossRef]
  85. Sfameni, S.; Truant, F.; Visco, A.; Plutino, M.R. Design and Development of Sustainable Nanotechnological Protective Systems in Surface Treatment. Ph.D. Thesis, University of Messina, Messina, Italy, 2022. Available online: https://hdl.handle.net/11570/3243553 (accessed on 11 February 2023).
  86. Cui, G.; Bi, Z.; Zhang, R.; Liu, J.; Yu, X.; Li, Z. A Comprehensive Review on Graphene-Based Anti-Corrosive Coatings. Chem. Eng. J. 2019, 373, 104–121. [Google Scholar] [CrossRef]
  87. Teng, S.; Gao, Y.; Cao, F.; Kong, D.; Zheng, X.; Ma, X.; Zhi, L. Zinc-Reduced Graphene Oxide for Enhanced Corrosion Protection of Zinc-Rich Epoxy Coatings. Prog. Org. Coat. 2018, 123, 185–189. [Google Scholar] [CrossRef]
  88. Bat, E.; Gündüz, G.; Kısakürek, D.; Akhmedov, İ.M. Synthesis and Characterization of Hyperbranched and Air Drying Fatty Acid Based Resins. Prog. Org. Coat. 2006, 55, 330–336. [Google Scholar] [CrossRef]
  89. Singh, A.P.; Suryanarayana, C.; Baloji Naik, R.; Gunasekaran, G. Development of Hyperbranched Polyester Polyol-Based Waterborne Anticorrosive Coating. J. Coat. Technol. Res. 2016, 13, 41–51. [Google Scholar] [CrossRef]
  90. Wang, G.; Wen, S.; Qian, S.; Wang, J.; Wang, C.; Chen, Y. Synthesis of Novel Nano Hyperbranched Polymer Resin and Its Corrosion Resistance in Coatings. Prog. Org. Coat. 2020, 140, 105496. [Google Scholar] [CrossRef]
  91. Payra, D.; Naito, M.; Fujii, Y.; Nagao, Y. Hydrophobized Plant Polyphenols: Self-Assembly and Promising Antibacterial, Adhesive, and Anticorrosion Coatings. Chem. Commun. 2016, 52, 312–315. [Google Scholar] [CrossRef] [PubMed]
  92. Dargahi, M.; Olsson, A.L.J.; Tufenkji, N.; Gaudreault, R. Green Technology: Tannin-Based Corrosion Inhibitor for Protection of Mild Steel. Corrosion 2015, 71, 1321–1329. [Google Scholar] [CrossRef] [PubMed]
  93. Khajali, F.; Slominski, B.A. Factors That Affect the Nutritive Value of Canola Meal for Poultry. Poult. Sci. 2012, 91, 2564–2575. [Google Scholar] [CrossRef]
  94. Nouailhas, H.; Aouf, C.; Le Guerneve, C.; Caillol, S.; Boutevin, B.; Fulcrand, H. Synthesis and Properties of Biobased Epoxy Resins. Part 1. Glycidylation of Flavonoids by Epichlorohydrin. J. Polym. Sci. Part A Polym. Chem. 2011, 49, 2261–2270. [Google Scholar] [CrossRef] [Green Version]
  95. Bloot, A.P.M.; Kalschne, D.L.; Amaral, J.A.S.; Baraldi, I.J.; Canan, C. A Review of Phytic Acid Sources, Obtention, and Applications. Food Rev. Int. 2021, 1–20. [Google Scholar] [CrossRef]
  96. Nita, L.E.; Chiriac, A.P.; Ghilan, A.; Rusu, A.G.; Tudorachi, N.; Timpu, D. Alginate Enriched with Phytic Acid for Hydrogels Preparation. Int. J. Biol. Macromol. 2021, 181, 561–571. [Google Scholar] [CrossRef] [PubMed]
  97. Chen, G.; Yuan, B.; Wang, Y.; Chen, X.; Huang, C.; Shang, S.; Tao, H.; Liu, J.; Sun, W.; Yang, P.; et al. Nacre-Biomimetic Graphene Oxide Paper Intercalated by Phytic Acid and Its Ultrafast Fire-Alarm Application. J. Colloid Interface Sci. 2020, 578, 412–421. [Google Scholar] [CrossRef]
  98. Handa, V.; Sharma, D.; Kaur, A.; Arya, S.K. Biotechnological Applications of Microbial Phytase and Phytic Acid in Food and Feed Industries. Biocatal. Agric. Biotechnol. 2020, 25, 101600. [Google Scholar] [CrossRef]
  99. Florea, R.M.; Carcea, I. Sustainable Anti-Corrosive Protection Technologies for Metal Products by Electrodeposition of HEA Layers. IOP Conf. Ser. Mater. Sci. Eng. 2019, 591, 12014. [Google Scholar] [CrossRef]
  100. Verma, C.; Ebenso, E.E.; Quraishi, M.A. Corrosion Inhibitors for Ferrous and Non-Ferrous Metals and Alloys in Ionic Sodium Chloride Solutions: A Review. J. Mol. Liq. 2017, 248, 927–942. [Google Scholar] [CrossRef]
  101. Visakh, P.M.; Yoshihiko, A. Flame Retardants: Polymer Blends, Composites and Nanocomposites; Springer: Berlin/Heidelberg, Germany, 2015; ISBN 3319034677. [Google Scholar]
  102. He, W.; Song, P.; Yu, B.; Fang, Z.; Wang, H. Flame Retardant Polymeric Nanocomposites through the Combination of Nanomaterials and Conventional Flame Retardants. Prog. Mater. Sci. 2020, 114, 100687. [Google Scholar] [CrossRef]
  103. Ielo, I.; Giacobello, F.; Sfameni, S.; Rando, G.; Galletta, M.; Trovato, V.; Rosace, G.; Plutino, M.R. Nanostructured Surface Finishing and Coatings: Functional Properties and Applications. Materials 2021, 14, 2733. [Google Scholar] [CrossRef] [PubMed]
  104. Vahabi, H.; Laoutid, F.; Mehrpouya, M.; Saeb, M.R.; Dubois, P. Flame Retardant Polymer Materials: An Update and the Future for 3D Printing Developments. Mater. Sci. Eng. R Rep. 2021, 144, 100604. [Google Scholar] [CrossRef]
  105. Zhao, S.; Chen, X.; Zhou, Y.; Zhao, B.; Hu, Q.; Chen, S.; Pan, K. Molecular Design of Reactive Flame Retardant for Preparing Biobased Flame Retardant Polyamide 56. Polym. Degrad. Stab. 2023, 207, 110212. [Google Scholar] [CrossRef]
  106. Alongi, J.; Ciobanu, M.; Malucelli, G. Sol–Gel Treatments for Enhancing Flame Retardancy and Thermal Stability of Cotton Fabrics: Optimisation of the Process and Evaluation of the Durability. Cellulose 2011, 18, 167–177. [Google Scholar] [CrossRef]
  107. Castellano, A.; Colleoni, C.; Iacono, G.; Mezzi, A.; Plutino, M.R.; Malucelli, G.; Rosace, G. Synthesis and Characterization of a Phosphorous/Nitrogen Based Sol-Gel Coating as a Novel Halogen- and Formaldehyde-Free Flame Retardant Finishing for Cotton Fabric. Polym. Degrad. Stab. 2019, 162, 148–159. [Google Scholar] [CrossRef]
  108. Vahabi, H.; Shabanian, M.; Aryanasab, F.; Mangin, R.; Laoutid, F.; Saeb, M.R. Inclusion of Modified Lignocellulose and Nano-Hydroxyapatite in Development of New Bio-Based Adjuvant Flame Retardant for Poly(Lactic Acid). Thermochim. Acta 2018, 666, 51–59. [Google Scholar] [CrossRef]
  109. Passauer, L. Thermal Characterization of Ammonium Starch Phosphate Carbamates for Potential Applications as Bio-Based Flame-Retardants. Carbohydr. Polym. 2019, 211, 69–74. [Google Scholar] [CrossRef]
  110. Alongi, J.; Carletto, R.A.; Bosco, F.; Carosio, F.; Di Blasio, A.; Cuttica, F.; Antonucci, V.; Giordano, M.; Malucelli, G. Caseins and Hydrophobins as Novel Green Flame Retardants for Cotton Fabrics. Polym. Degrad. Stab. 2014, 99, 111–117. [Google Scholar] [CrossRef]
  111. Basak, S.; Ali, S.W. Sustainable Fire Retardancy of Textiles Using Bio-Macromolecules. Polym. Degrad. Stab. 2016, 133, 47–64. [Google Scholar] [CrossRef]
  112. Danyliuk, N.; Tomaszewska, J.; Tatarchuk, T. Halloysite Nanotubes and Halloysite-Based Composites for Environmental and Biomedical Applications. J. Mol. Liq. 2020, 309, 113077. [Google Scholar] [CrossRef]
  113. Otitoju, T.A.; Ahmad, A.L.; Ooi, B.S. Recent Advances in Hydrophilic Modification and Performance of Polyethersulfone (PES) Membrane via Additive Blending. RSC Adv. 2018, 8, 22710–22728. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. He, Y.; Xu, W.; Tang, R.; Zhang, C.; Yang, Q. PH-Responsive Nanovalves Based on Encapsulated Halloysite for the Controlled Release of a Corrosion Inhibitor in Epoxy Coating. RSC Adv. 2015, 5, 90609–90620. [Google Scholar] [CrossRef]
  115. Goda, E.S.; Yoon, K.R.; El-sayed, S.H.; Hong, S.E. Halloysite Nanotubes as Smart Flame Retardant and Economic Reinforcing Materials: A Review. Thermochim. Acta 2018, 669, 173–184. [Google Scholar] [CrossRef]
  116. Rando, G.; Sfameni, S.; Plutino, M.R. Development of Functional Hybrid Polymers and Gel Materials for Sustainable Membrane-Based Water Treatment Technology: How to Combine Greener and Cleaner Approaches. Gels 2023, 9, 9. [Google Scholar] [CrossRef] [PubMed]
  117. Boddula, R.; Ahamed, M.I.; Asiri, A.M. Polymers Coatings: Technology and Applications; John Wiley & Sons: Hoboken, NJ, USA, 2020; ISBN 1119654998. [Google Scholar]
  118. Cirisano, F.; Ferrari, M. Sustainable Materials for Liquid Repellent Coatings. Coatings 2021, 11, 1508. [Google Scholar] [CrossRef]
  119. PV, V.M.; Kudapa, V.K. Recent Developments in Usage of Fluorine-Free Nano Structured Materials in Oil-Water Separation: A Review. Surf. Interfaces 2021, 27, 101455. [Google Scholar] [CrossRef]
  120. Yang, Z.; Liu, X.; Tian, Y. Fabrication of Super-Hydrophobic Nickel Film on Copper Substrate with Improved Corrosion Inhibition by Electrodeposition Process. Colloids Surf. A Physicochem. Eng. Asp. 2019, 560, 205–212. [Google Scholar] [CrossRef] [Green Version]
  121. Deeksha, P.; Deepika, G.; Nishanthini, J.; Hikku, G.S.; Jeyasubramanian, K.; Murugesan, R. Super-Hydrophobicity: Mechanism, Fabrication and Its Application in Medical Implants to Prevent Biomaterial Associated Infections. J. Ind. Eng. Chem. 2020, 92, 1–17. [Google Scholar] [CrossRef]
  122. Sznajder-Katarzyńska, K.; Surma, M.; Cieślik, I. A Review of Perfluoroalkyl Acids (PFAAs) in Terms of Sources, Applications, Human Exposure, Dietary Intake, Toxicity, Legal Regulation, and Methods of Determination. J. Chem. 2019, 2019, 2717528. [Google Scholar] [CrossRef] [Green Version]
  123. Prontera, C.T.; Sico, G.; Montanino, M.; De Girolamo Del Mauro, A.; Tassini, P.; Maglione, M.G.; Minarini, C.; Manini, P. Sustainable, Fluorine-Free, Low Cost and Easily Processable Materials for Hydrophobic Coatings on Flexible Plastic Substrates. Materials 2019, 12, 2234. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Giacobello, F.; Ielo, I.; Belhamdi, H.; Plutino, M.R. Geopolymers and Functionalization Strategies for the Development of Sustainable Materials in Construction Industry and Cultural Heritage Applications: A Review. Materials 2022, 15, 1725. [Google Scholar] [CrossRef] [PubMed]
  125. Camuffo, D. Microclimate for Cultural Heritage: Measurement, Risk Assessment, Conservation, Restoration, and Maintenance of Indoor and Outdoor Monuments; Elsevier: Amsterdam, The Netherlands, 2019; ISBN 0444641076. [Google Scholar]
  126. Trovato, V.; Rosace, G.; Colleoni, C.; Sfameni, S.; Migani, V.; Plutino, M.R. Sol-Gel Based Coatings for the Protection of Cultural Heritage Textiles. IOP Conf. Ser. Mater. Sci. Eng. 2020, 777, 12007. [Google Scholar] [CrossRef]
  127. Fistos, T.; Fierascu, I.; Doni, M.; Chican, I.E.; Fierascu, R.C. A Short Overview of Recent Developments in the Application of Polymeric Materials for the Conservation of Stone Cultural Heritage Elements. Materials 2022, 15, 6294. [Google Scholar] [CrossRef]
  128. Wang, W.; Meng, Q.; Li, Q.; Liu, J.; Zhou, M.; Jin, Z.; Zhao, K. Chitosan Derivatives and Their Application in Biomedicine. Int. J. Mol. Sci. 2020, 21, 487. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Shahid-Ul-Islam; Shahid, M.; Mohammad, F. Green Chemistry Approaches to Develop Antimicrobial Textiles Based on Sustainable Biopolymers—A Review. Ind. Eng. Chem. Res. 2013, 52, 5245–5260. [Google Scholar] [CrossRef]
  130. Galiano, F.; Briceño, K.; Marino, T.; Molino, A.; Christensen, K.V.; Figoli, A. Advances in Biopolymer-Based Membrane Preparation and Applications. J. Memb. Sci. 2018, 564, 562–586. [Google Scholar] [CrossRef]
  131. Hu, J. Active Coatings for Smart Textiles; Woodhead Publishing: Sawston, UK, 2016; ISBN 0081002653. [Google Scholar]
  132. Ramasubramanian, B.; Sundarrajan, S.; Rao, R.P.; Reddy, M.V.; Chellappan, V.; Ramakrishna, S. Novel Low-Carbon Energy Solutions for Powering Emerging Wearables, Smart Textiles, and Medical Devices. Energy Environ. Sci. 2022, 15, 4928–4981. [Google Scholar] [CrossRef]
  133. Montazer, M.; Harifi, T. Nanofinishing of Textile Materials; Woodhead Publishing: Sawston, UK, 2018; ISBN 0081012500. [Google Scholar]
  134. AbouElmaaty, T.; Abdeldayem, S.A.; Ramadan, S.M.; Sayed-Ahmed, K.; Plutino, M.R. Coloration and Multi-Functionalization of Polypropylene Fabrics with Selenium Nanoparticles. Polymers 2021, 13, 2483. [Google Scholar] [CrossRef]
  135. Smith, W.C. Smart Textile Coatings and Laminates; Elsevier: Amsterdam, The Netherlands, 2010; ISBN 1845697782. [Google Scholar]
  136. Trovato, V.; Mezzi, A.; Brucale, M.; Abdeh, H.; Drommi, D.; Rosace, G.; Plutino, M.R. Sol-Gel Assisted Immobilization of Alizarin Red S on Polyester Fabrics for Developing Stimuli-Responsive Wearable Sensors. Polymers 2022, 14, 2788. [Google Scholar] [CrossRef]
  137. Jones, F.N.; Nichols, M.E.; Pappas, S.P. Organic Coatings: Science and Technology; John Wiley & Sons: Hoboken, NJ, USA, 2017; ISBN 111902689X. [Google Scholar]
  138. Müller, B.; Poth, U. Coatings Formulation. In Coatings Formulation; Vincentz Network: Hanover, Germany, 2017; ISBN 3748600267. [Google Scholar]
  139. Wang, S.; Xu, D.; Liu, Y.; Jiang, Z.; Zhu, P. Preparation and Mechanism of Phosphoramidate-Based Sol-Gel Coating for Flame-.Retardant Viscose Fabric. Polym. Degrad. Stab. 2021, 190, 109620. [Google Scholar] [CrossRef]
  140. Malucelli, G. Sol-Gel and Layer-by-Layer Coatings for Flame-Retardant Cotton Fabrics: Recent Advances. Coatings 2020, 10, 333. [Google Scholar] [CrossRef] [Green Version]
  141. Puoci, F.; Saturnino, C.; Trovato, V.; Iacopetta, D.; Piperopoulos, E.; Triolo, C.; Bonomo, M.G.; Drommi, D.; Parisi, O.I.; Milone, C.; et al. Sol-Gel Treatment of Textiles for the Entrapping of an Antioxidant/Anti-Inflammatory Molecule: Functional Coating Morphological Characterization and Drug Release Evaluation. Appl. Sci. 2020, 10, 2287. [Google Scholar] [CrossRef] [Green Version]
  142. Camlibel, N.O.; Arik, B. Sol–Gel Applications in Textile Finishing Processes. In Recent Applications in Sol-Gel Synthesis; IntechOpen Limited: London, UK, 2017; pp. 253–281. [Google Scholar]
  143. Abidi, N.; Hequet, E.; Tarimala, S.; Dai, L.L. Cotton Fabric Surface Modification for Improved UV Radiation Protection Using Sol–Gel Process. J. Appl. Polym. Sci. 2007, 104, 111–117. [Google Scholar] [CrossRef]
  144. Mahltig, B.; Textor, T. Combination of Silica Sol and Dyes on Textiles. J. Sol-Gel Sci. Technol. 2006, 39, 111–118. [Google Scholar] [CrossRef]
  145. Poli, R.; Colleoni, C.; Calvimontes, A.; Polášková, H.; Dutschk, V.; Rosace, G. Innovative Sol–Gel Route in Neutral Hydroalcoholic Condition to Obtain Antibacterial Cotton Finishing by Zinc Precursor. J. Sol-Gel Sci. Technol. 2015, 74, 151–160. [Google Scholar] [CrossRef]
  146. Periyasamy, A.P.; Venkataraman, M.; Kremenakova, D.; Militky, J.; Zhou, Y. Progress in Sol-Gel Technology for the Coatings of Fabrics. Materials 2020, 13, 1838. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Libertino, S.; Plutino, M.R.; Rosace, G. Design and Development of Wearable Sensing Nanomaterials for Smart Textiles. AIP Conf. Proc. 2018, 1990, 020016. [Google Scholar] [CrossRef]
  148. Kappes, R.S.; Urbainczyk, T.; Artz, U.; Textor, T.; Gutmann, J.S. Flame Retardants Based on Amino Silanes and Phenylphosphonic Acid. Polym. Degrad. Stab. 2016, 129, 168–179. [Google Scholar] [CrossRef]
  149. Sanchez, C.; Julián, B.; Belleville, P.; Popall, M. Applications of Hybrid Organic–Inorganic Nanocomposites. J. Mater. Chem. 2005, 15, 3559. [Google Scholar] [CrossRef]
  150. Deng, Y.; Peng, C.; Dai, M.; Lin, D.; Ali, I.; Alhewairini, S.S.; Zheng, X.; Chen, G.; Li, J.; Naz, I. Recent Development of Super-Wettable Materials and Their Applications in Oil-Water Separation. J. Clean. Prod. 2020, 266, 121624. [Google Scholar] [CrossRef]
  151. Zeng, Q.; Zhou, H.; Huang, J.; Guo, Z. Review on the Recent Development of Durable Superhydrophobic Materials for Practical Applications. Nanoscale 2021, 13, 11734–11764. [Google Scholar] [CrossRef]
  152. Amiri, S.; Rahimi, A. Hybrid Nanocomposite Coating by Sol–Gel Method: A Review. Iran. Polym. J. 2016, 25, 559–577. [Google Scholar] [CrossRef]
  153. Zhou, H.; Wang, H.; Niu, H.; Gestos, A.; Lin, T. Robust, Self-Healing Superamphiphobic Fabrics Prepared by Two-Step Coating of Fluoro-Containing Polymer, Fluoroalkyl Silane, and Modified Silica Nanoparticles. Adv. Funct. Mater. 2013, 23, 1664–1670. [Google Scholar] [CrossRef]
  154. Ferrero, F.; Periolatto, M. Application of Fluorinated Compounds to Cotton Fabrics via Sol–Gel. Appl. Surf. Sci. 2013, 275, 201–207. [Google Scholar] [CrossRef]
  155. Wang, Z.; Cousins, I.T.; Scheringer, M.; Hungerbühler, K. Fluorinated Alternatives to Long-Chain Perfluoroalkyl Carboxylic Acids (PFCAs), Perfluoroalkane Sulfonic Acids (PFSAs) and Their Potential Precursors. Environ. Int. 2013, 60, 242–248. [Google Scholar] [CrossRef]
  156. Yu, C.; Shi, K.; Ning, J.; Zheng, Z.; Yu, H.; Yang, Z.; Liu, J. Preparation and Application of Fluorine-Free Finishing Agent with Excellent Water Repellency for Cotton Fabric. Polymers 2021, 13, 2980. [Google Scholar] [CrossRef]
  157. Lei, H.; Xiong, M.; Xiao, J.; Zheng, L.; Zhu, Y.; Li, X.; Zhuang, Q.; Han, Z. Fluorine-Free Low Surface Energy Organic Coating for Anti-Stain Applications. Prog. Org. Coat. 2017, 103, 182–192. [Google Scholar] [CrossRef]
  158. Sfameni, S.; Lawnick, T.; Rando, G.; Visco, A.; Textor, T.; Plutino, M.R. Super-Hydrophobicity of Polyester Fabrics Driven by Functional Sustainable Fluorine-Free Silane-Based Coatings. Gels 2023, 9, 109. [Google Scholar] [CrossRef] [PubMed]
  159. Das, A.; Mahanwar, P. A Brief Discussion on Advances in Polyurethane Applications. Adv. Ind. Eng. Polym. Res. 2020, 3, 93–101. [Google Scholar] [CrossRef]
  160. Si, P.; Zhao, B. Water-Based Polyurethanes for Sustainable Advanced Manufacture. Can. J. Chem. Eng. 2021, 99, 1851–1869. [Google Scholar] [CrossRef]
  161. Li, Y.; Ning, C. Latest Research Progress of Marine Microbiological Corrosion and Bio-Fouling, and New Approaches of Marine Anti-Corrosion and Anti-Fouling. Bioact. Mater. 2019, 4, 189–195. [Google Scholar] [CrossRef] [PubMed]
  162. Zhang, X.; Brodus, D.; Hollimon, V.; Hu, H. A Brief Review of Recent Developments in the Designs That Prevent Bio-Fouling on Silicon and Silicon-Based Materials. Chem. Cent. J. 2017, 11, 18. [Google Scholar] [CrossRef] [Green Version]
  163. Hakim, M.L.; Nugroho, B.; Nurrohman, M.N.; Suastika, I.K.; Utama, I.K.A.P. Investigation of Fuel Consumption on an Operating Ship Due to Biofouling Growth and Quality of Anti-Fouling Coating. IOP Conf. Ser. Earth Environ. Sci. 2019, 339, 12037. [Google Scholar] [CrossRef]
  164. Selim, M.S.; Shenashen, M.A.; El-Safty, S.A.; Higazy, S.A.; Selim, M.M.; Isago, H.; Elmarakbi, A. Recent Progress in Marine Foul-Release Polymeric Nanocomposite Coatings. Prog. Mater. Sci. 2017, 87, 1–32. [Google Scholar] [CrossRef]
  165. Ali, A.; Jamil, M.I.; Jiang, J.; Shoaib, M.; Amin, B.U.; Luo, S.; Zhan, X.; Chen, F.; Zhang, Q. An Overview of Controlled-Biocide-Release Coating Based on Polymer Resin for Marine Antifouling Applications. J. Polym. Res. 2020, 27, 85. [Google Scholar] [CrossRef]
  166. Simovich, T.; Rosenhahn, A.; Lamb, R.N. Thermoregeneration of Plastrons on Superhydrophobic Coatings for Sustained Antifouling Properties. Adv. Eng. Mater. 2020, 22, 1900806. [Google Scholar] [CrossRef]
  167. Ielo, I.; Giacobello, F.; Castellano, A.; Sfameni, S.; Rando, G.; Plutino, M.R. Development of Antibacterial and Antifouling Innovative and Eco-Sustainable Sol–Gel Based Materials: From Marine Areas Protection to Healthcare Applications. Gels 2022, 8, 26. [Google Scholar]
  168. Maan, A.M.C.; Hofman, A.H.; de Vos, W.M.; Kamperman, M. Recent Developments and Practical Feasibility of Polymer-Based Antifouling Coatings. Adv. Funct. Mater. 2020, 30, 3–8. [Google Scholar] [CrossRef]
  169. Nurioglu, A.G.; Esteves, A.C.C.; De With, G. Non-Toxic, Non-Biocide-Release Antifouling Coatings Based on Molecular Structure Design for Marine Applications. J. Mater. Chem. B 2015, 3, 6547–6570. [Google Scholar] [CrossRef] [Green Version]
  170. Rosace, G.; Cardiano, P.; Urzì, C.; De Leo, F.; Galletta, M.; Ielo, I.; Plutino, M.R. Potential Roles of Fluorine-Containing Sol-Gel Coatings against Adhesion to Control Microbial Biofilm. IOP Conf. Ser. Mater. Sci. Eng. 2018, 459, 12021. [Google Scholar] [CrossRef]
  171. Lazzara, G.; Cavallaro, G.; Panchal, A.; Fakhrullin, R.; Stavitskaya, A.; Vinokurov, V.; Lvov, Y. An Assembly of Organic-Inorganic Composites Using Halloysite Clay Nanotubes. Curr. Opin. Colloid Interface Sci. 2018, 35, 42–50. [Google Scholar] [CrossRef]
  172. Yah, W.O.; Takahara, A.; Lvov, Y.M. Selective Modification of Halloysite Lumen with Octadecylphosphonic Acid: New Inorganic Tubular Micelle. J. Am. Chem. Soc. 2012, 134, 1853–1859. [Google Scholar] [CrossRef]
  173. Khatoon, N.; Chu, M.Q.; Zhou, C.H. Nanoclay-Based Drug Delivery Systems and Their Therapeutic Potentials. J. Mater. Chem. B 2020, 8, 7335–7351. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Functional and sustainable coating features and their potential application sectors.
Figure 1. Functional and sustainable coating features and their potential application sectors.
Ijms 24 05472 g001
Figure 2. Hydrolysis and condensation reaction steps of silane alkoxide precursors occurring in acidic (a,b) and alkaline environments (c,d). Reproduced from MDPI Ref. [31].
Figure 2. Hydrolysis and condensation reaction steps of silane alkoxide precursors occurring in acidic (a,b) and alkaline environments (c,d). Reproduced from MDPI Ref. [31].
Ijms 24 05472 g002
Figure 3. Recent trends in sustainable anticorrosive coating production. Reproduced from Ref. [85].
Figure 3. Recent trends in sustainable anticorrosive coating production. Reproduced from Ref. [85].
Ijms 24 05472 g003
Figure 4. Graphene oxide intercalated phytic acid interaction (a) and stability of the obtained water dispersions (b). Sol–gel-based approach with proper silane precursors for the production of sustainable waterborne anticorrosive protective coatings (c) and cross-linked matrix (d). Reproduced from MDPI Ref. [57].
Figure 4. Graphene oxide intercalated phytic acid interaction (a) and stability of the obtained water dispersions (b). Sol–gel-based approach with proper silane precursors for the production of sustainable waterborne anticorrosive protective coatings (c) and cross-linked matrix (d). Reproduced from MDPI Ref. [57].
Ijms 24 05472 g004
Figure 5. Self-healing mechanism of tannin-functionalized coating mixed with silane- hybrid ZnO nanoparticles, with passivation of the damaged steel surface by the formation of tannates. Reprinted with permission from Ref. [59]. Copyright 2019, A.F. Jaramillo, Progress in Organic Coatings, Elsevier.
Figure 5. Self-healing mechanism of tannin-functionalized coating mixed with silane- hybrid ZnO nanoparticles, with passivation of the damaged steel surface by the formation of tannates. Reprinted with permission from Ref. [59]. Copyright 2019, A.F. Jaramillo, Progress in Organic Coatings, Elsevier.
Ijms 24 05472 g005
Figure 6. Preparation of the hybrid epoxy coating containing precipitated amorphous silica functionalized with MIL extract with its anticorrosive behavior (a) and mechanism of corrosion inhibition of bare commercial steel and MIL-adsorbed commercial steel in a saline environment (b). Adapted from Ref. [60]. Copyright 2019, K. K. Veedu, ACS Omega, American Chemical Society.
Figure 6. Preparation of the hybrid epoxy coating containing precipitated amorphous silica functionalized with MIL extract with its anticorrosive behavior (a) and mechanism of corrosion inhibition of bare commercial steel and MIL-adsorbed commercial steel in a saline environment (b). Adapted from Ref. [60]. Copyright 2019, K. K. Veedu, ACS Omega, American Chemical Society.
Ijms 24 05472 g006
Figure 7. Representation of the combustion process of polymeric materials. Reprinted with permission from Ref. [102]. Copyright 2020, W. He, Progress in Materials Science, Elsevier.
Figure 7. Representation of the combustion process of polymeric materials. Reprinted with permission from Ref. [102]. Copyright 2020, W. He, Progress in Materials Science, Elsevier.
Ijms 24 05472 g007
Figure 8. Different agents employed for the preparation of flame-retardant polymeric composites from nanomaterials (a) to bio-based derivatives (b). Reprinted with permission from Ref. [102]; copyright 2020, W. He, Progress in Materials Science, Elsevier and Ref. [104] copyright 2021, H. Vahabi, Materials Science and Engineering: R: Reports, Elsevier, respectively.
Figure 8. Different agents employed for the preparation of flame-retardant polymeric composites from nanomaterials (a) to bio-based derivatives (b). Reprinted with permission from Ref. [102]; copyright 2020, W. He, Progress in Materials Science, Elsevier and Ref. [104] copyright 2021, H. Vahabi, Materials Science and Engineering: R: Reports, Elsevier, respectively.
Ijms 24 05472 g008
Figure 9. Possible mechanism of flame-retardant action of the eco-friendly composite film of epoxy resin containing spent coffee waste treated with phosphorous. Reproduced from MDPI Ref. [61].
Figure 9. Possible mechanism of flame-retardant action of the eco-friendly composite film of epoxy resin containing spent coffee waste treated with phosphorous. Reproduced from MDPI Ref. [61].
Ijms 24 05472 g009
Figure 10. Stone weathering phenomena (a) and protective features and implemented properties introduced by the employing of functional polymeric formulations as coatings (b). Reproduced from MDPI Ref. [127].
Figure 10. Stone weathering phenomena (a) and protective features and implemented properties introduced by the employing of functional polymeric formulations as coatings (b). Reproduced from MDPI Ref. [127].
Ijms 24 05472 g010
Figure 11. Schematization of the formation of the hydrophobic zein coating on Serena stone by the spray-coating technique. Reprinted with permission from Ref. [65]. Copyright 2021, M. Zucchelli, Progress in Organic Coatings, Elsevier.
Figure 11. Schematization of the formation of the hydrophobic zein coating on Serena stone by the spray-coating technique. Reprinted with permission from Ref. [65]. Copyright 2021, M. Zucchelli, Progress in Organic Coatings, Elsevier.
Ijms 24 05472 g011
Figure 12. Some application fields of functional and coted textile fabrics. Reproduced from MDPI Ref. [67].
Figure 12. Some application fields of functional and coted textile fabrics. Reproduced from MDPI Ref. [67].
Ijms 24 05472 g012
Figure 13. Schematic approach of sol–gel dip-padding to functionalize textile fabrics. Reproduced from MDPI Ref. [34].
Figure 13. Schematic approach of sol–gel dip-padding to functionalize textile fabrics. Reproduced from MDPI Ref. [34].
Ijms 24 05472 g013
Figure 14. Cotton fabrics are functionalized in two steps with an alkyl(trialkoxy)silane polymer shell using covalent “grafting to” or “grafting from” procedures (a). The coated cotton fabrics’ surface hydrophobicity (b). Reproduced from MDPI Ref. [67].
Figure 14. Cotton fabrics are functionalized in two steps with an alkyl(trialkoxy)silane polymer shell using covalent “grafting to” or “grafting from” procedures (a). The coated cotton fabrics’ surface hydrophobicity (b). Reproduced from MDPI Ref. [67].
Ijms 24 05472 g014
Figure 15. The condensation process between the cellulose and alkoxysilane ends (a), followed by the anchoring of alkyl(trialkoxy)silanes (b). Reproduced from MDPI Ref. [67].
Figure 15. The condensation process between the cellulose and alkoxysilane ends (a), followed by the anchoring of alkyl(trialkoxy)silanes (b). Reproduced from MDPI Ref. [67].
Ijms 24 05472 g015
Figure 16. Schematic preparation of the SFP from CTA and PEG200 (a) and subsequent synthesis of the waterborne polyurethane (b) (MDI = 4,4′-methylenebisphenyl diisocyanate; PCL = polycaprolactone diol 2000; TEA = triethylamine). Reprinted with permission Ref. [69]. Copyright 2019, I. Bramhecha, Ind. Eng. Chem. Res., American Chemical Society.
Figure 16. Schematic preparation of the SFP from CTA and PEG200 (a) and subsequent synthesis of the waterborne polyurethane (b) (MDI = 4,4′-methylenebisphenyl diisocyanate; PCL = polycaprolactone diol 2000; TEA = triethylamine). Reprinted with permission Ref. [69]. Copyright 2019, I. Bramhecha, Ind. Eng. Chem. Res., American Chemical Society.
Ijms 24 05472 g016
Figure 17. The three main antifouling strategies are shown schematically as preventing foulants from adhering to the surface (fouling-resistant) (a), reducing foulant interaction with the surface (fouling-release) (b), and degrading or eliminating bio-foulants (fouling-degrading) (c). Methods for giving a surface antifouling characteristics, including changing the surface chemistry (d), changing the surface topography (e), changing the coating design (f). Reproduced from MDPI Ref. [167].
Figure 17. The three main antifouling strategies are shown schematically as preventing foulants from adhering to the surface (fouling-resistant) (a), reducing foulant interaction with the surface (fouling-release) (b), and degrading or eliminating bio-foulants (fouling-degrading) (c). Methods for giving a surface antifouling characteristics, including changing the surface chemistry (d), changing the surface topography (e), changing the coating design (f). Reproduced from MDPI Ref. [167].
Ijms 24 05472 g017
Figure 18. Antifouling and fouling-release activity of biocide-release and non-biocide-release coatings, respectively. Reproduced from MDPI Ref. [70].
Figure 18. Antifouling and fouling-release activity of biocide-release and non-biocide-release coatings, respectively. Reproduced from MDPI Ref. [70].
Ijms 24 05472 g018
Figure 19. Synthetic steps involving the production of the eco-friendly sol–gel hydrophobic antifouling coating. Reproduced from MDPI Ref. [71].
Figure 19. Synthetic steps involving the production of the eco-friendly sol–gel hydrophobic antifouling coating. Reproduced from MDPI Ref. [71].
Ijms 24 05472 g019
Figure 20. Curcumin-based benzoxazine monomer synthesis reaction (a) and synthetic pathway for the curcumin-based benzoxazine polymerization through thermal induction of ring opening (b). Reprinted with permission from Ref. [73]. Copyright 2021, Y. Deng, Composites Part B: Engineering, Elsevier.
Figure 20. Curcumin-based benzoxazine monomer synthesis reaction (a) and synthetic pathway for the curcumin-based benzoxazine polymerization through thermal induction of ring opening (b). Reprinted with permission from Ref. [73]. Copyright 2021, Y. Deng, Composites Part B: Engineering, Elsevier.
Ijms 24 05472 g020
Figure 21. Arbutin-based benzoxazine monomer synthesis reaction (a) and synthetic pathway for the arbutin-based benzoxazine polymerization (b). Reprinted with permission from Ref. [74]. Copyright 2022, T. Periyasamy, Progress in Organic Coatings, Elsevier.
Figure 21. Arbutin-based benzoxazine monomer synthesis reaction (a) and synthetic pathway for the arbutin-based benzoxazine polymerization (b). Reprinted with permission from Ref. [74]. Copyright 2022, T. Periyasamy, Progress in Organic Coatings, Elsevier.
Ijms 24 05472 g021
Figure 22. Synthetic pathway of polyimide (a) by liquid flame spray approach (b). Reprinted with permission from Ref. [75]. Copyright 2017, Y. Liu, Materials & Design, Elsevier.
Figure 22. Synthetic pathway of polyimide (a) by liquid flame spray approach (b). Reprinted with permission from Ref. [75]. Copyright 2017, Y. Liu, Materials & Design, Elsevier.
Ijms 24 05472 g022
Figure 23. Schematic preparation of the eco-friendly sodium salicylate and N-(2,4,6-trichlorophenyl)maleimide biocide agent-loaded HNTs (a) and antifouling behavior of the hybrid composite epoxy-based paint obtained from the incorporation of the functional HNTs (b). Reprinted with permission from Ref. [77]. Copyright 2021, M. Tonelli, Colloids and Surfaces A: Physicochemical and Engineering Aspects, Elsevier.
Figure 23. Schematic preparation of the eco-friendly sodium salicylate and N-(2,4,6-trichlorophenyl)maleimide biocide agent-loaded HNTs (a) and antifouling behavior of the hybrid composite epoxy-based paint obtained from the incorporation of the functional HNTs (b). Reprinted with permission from Ref. [77]. Copyright 2021, M. Tonelli, Colloids and Surfaces A: Physicochemical and Engineering Aspects, Elsevier.
Ijms 24 05472 g023
Table 1. Some recent developments in the design and application of functional and sustainable coatings based on natural substances, waste-derived compounds, waterborne formulations and eco-friendly approaches.
Table 1. Some recent developments in the design and application of functional and sustainable coatings based on natural substances, waste-derived compounds, waterborne formulations and eco-friendly approaches.
Functional PrecursorsDeposition ApproachCoated SurfacePropertiesRef.
GPTMS 1, APTES 2, GO 3 and phytic acidFilm coatingAQ-36 aluminum and QD-36 carbon steelAnticorrosive[57]
Polyaniline and reduced GOElectrodeposition5083 Al alloyAnticorrosive[58]
ZnO nanoparticles, APTES, Pinus radiata-derived tannin and epoxy resinSpray coatingASTM A36 steel platesHydrophobic, anticorrosive, self-healing[59]
Mangifera indica L. leaf extract, amorphous silica and epoxy resinDip-coatingCarbon steel Anticorrosive[60]
Phosphite-treated coffee biowaste and epoxy resinFilm-castingGlass substrate to peel off the final filmFlame retardant[61]
Halloysite and rennet caseinDip-pad-dry-cureCotton, polyester and blend of cotton and polyesterFlame retardant, antibacterial, antiviral[62]
Methyl methacrylate, TMSM 4, TEOS 5, FOTCS 6 and TiO2ImpregnationStone of a historical monumentThermal, mechanical and weathering resistance, hydrophobicity and self-cleaning[63]
Silane/siloxane emulsion, chitosan and AgNO3Spray-coating with airbrushBuilding porous limestoneWater repellent and biocide[64]
Zein solutionSpray coatingSandstoneWater repellent[65]
TEOS and citric acidSpray coating/dry-curing90% cotton and 10% polyester fabricsHydrophobicity[66]
GPTMS, Hexadecyltrimethoxysilane, Triethoxy(octyl)silane and Triethoxy(ethyl)silaneDip-pad-dry-cureCotton fabricsWater repellent, water-based anti-stain[67]
SWCNT 7 and WPUD 8Knife-coatingPolyester fabricsHydrophobic and conductive[68]
SFP 9, MDI 10 and PCL 11Knife on roller coatingCotton fabricsAntibacterial and breathable waterproof[69]
GPTMS, APTES, F3 12 and F16 13Film coating Glass platesHydrophobic and fouling-release[70]
GPTMS and alkyl(trialkoxy)silanesFilm coatingGlass plates treated with a commercial primer and tie-coatHydrophobic and fouling-release[71]
Fe-myristic acid compoundElectrodepositionCu metalSuperhydrophobic, bio-corrosion and bio-adhesion inhibition[72]
Curcumin, APTES, paraformaldehyde and PEG 14Film coatingCarbon steelAnticorrosive, inhibitor of microorganisms’ settlement, adhesion and growth[73]
Arbutin, AEAPTMS 15, paraformaldehyde and PEGDip coatingLow carbon steelAnticorrosive and anti-biofouling adhesion[74]
PMDA 16, ODA 17, Cu nanoparticlesLiquid flame sprayCarbon steel Q235Biocide-release antifouling[75]
Econea, diisocyanate and PDMS 18Dip-coatingAcrylic prototypesAntibacterial and bacteriostatic activity[76]
Halloysite, sodium salicylate, N-(2,4,6-trichlorophenyl)maleimide and epoxy paintFilm coatingPlastic sheetsBiocide-release antifouling[77]
1 GPTMS = (3-Glycidyloxypropyl)trimethoxysilane; 2 APTES = 3-(aminopropyl)triethoxysilane; 3 GO = graphene oxide; 4 TMSM = 3-(trimethoxysilyl)propyl methacrylate; 5 TEOS = tetraethyl orthosilicate; 6 FOCTS = perfluorooctyl-trichlorosilane; 7 SWCNT = single-walled carbon nanotubes; 8 WPUD = waterborne polyurethane-urea dispersions; 9 SFP = polyol made from citric acid and polyethylene glycol; 10 MDI = 4,4′-methylenebisphenyl diisocyanate; 11 PCL = polycaprolactone diol 2000; 12 F3 = 3,3,3-trifluoropropyl-trimethoxysilane; 13 F16 = glycidyl-2,2,3,3,4,4,5,5,6,6,7,7,8,8,9,9-hexadecafluorononylether; 14 PEG = polyethylene glycol; 15 AEAPTMS = [3-(2-aminoethylamino) propyl]trimethoxysilane; 16 PMDA = pyromellitic dianhydride; 17 ODA = 4,4′-oxydianiline; 18 PDMS = polydimethylsiloxane.
Table 2. Advantages and disadvantages of all reported strategies for sustainable and functional coating development.
Table 2. Advantages and disadvantages of all reported strategies for sustainable and functional coating development.
Coating StrategyAdvantagesDisadvantages
Sol–gel
  • High durability
  • Easy functionalization
  • Easy management
  • No cytotoxicity
  • Versatility
  • Uniform coating thickness
  • In some cases, needs of thermal curing
  • Hard to obtain crack-free coatings
  • Some substrates needs a pretreatment before the sol–gel deposition
  • Low porosity
(Blended)polymers and resins
  • Reduction of the fossil-based polymer content
  • High-mechanical features
  • Usage of a minimum of fossil-based epoxy resins and polymers
Bio-based systems
  • Zero fossil-based content
  • Tunable functionalities
  • Low processing temperatures
  • Starting materials and polymers have an unfavorable cost influence considering vast application scales
  • Lower durability related to their predisposition to biodegradation
Secondary-raw materials
  • Recycle of waste materials
  • Development of value-added coatings
  • Lower control of the homogeneity of the coating
  • Sometimes long synthetic protocols for the treatment of the waste matrix
Eco-friendly nanomaterials
  • Reduced use of toxic organic solvents
  • Improved active surface
  • Modulated functionalities
  • Often one-step synthesis protocols
  • Need for a proper design for the integration and compatibilization with the polymeric or organic/inorganic matrices
  • Requiring a strong immobilization in the matrix to avoid their release
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Sfameni, S.; Rando, G.; Plutino, M.R. Sustainable Secondary-Raw Materials, Natural Substances and Eco-Friendly Nanomaterial-Based Approaches for Improved Surface Performances: An Overview of What They Are and How They Work. Int. J. Mol. Sci. 2023, 24, 5472. https://doi.org/10.3390/ijms24065472

AMA Style

Sfameni S, Rando G, Plutino MR. Sustainable Secondary-Raw Materials, Natural Substances and Eco-Friendly Nanomaterial-Based Approaches for Improved Surface Performances: An Overview of What They Are and How They Work. International Journal of Molecular Sciences. 2023; 24(6):5472. https://doi.org/10.3390/ijms24065472

Chicago/Turabian Style

Sfameni, Silvia, Giulia Rando, and Maria Rosaria Plutino. 2023. "Sustainable Secondary-Raw Materials, Natural Substances and Eco-Friendly Nanomaterial-Based Approaches for Improved Surface Performances: An Overview of What They Are and How They Work" International Journal of Molecular Sciences 24, no. 6: 5472. https://doi.org/10.3390/ijms24065472

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop