Front. Mater. Sci. 2022, 16(3): 220607 https://doi.org/10.1007/s11706-022-0607-7 REVIEW ARTICLE Nanoparticles embedded into glass matrices: glass nanocomposites Javier FONSECA (✉) Department of Chemical Engineering, Northeastern University, Boston 02115, USA © Higher Education Press 2022 ABSTRACT: Research on glass nanocomposites (GNCs) has been very active in the past decades. GNCs have attracted — and still do — great interest in the fields of optoelectronics, photonics, sensing, electrochemistry, catalysis, biomedicine, and art. In this review, the potential applications of GNCs in these fields are briefly described to show the reader the possibilities of these materials. The most important synthesis methods of GNCs (melt-quenching, sol-gel, ion implantation, ion-exchange, staining process, spark plasma sintering, radio frequency sputtering, spray pyrolysis, and chemical vapor deposition techniques) are extensively explained. The major aim of this review is to systematize our knowledge about the synthesis of GNCs and to explore the mechanisms of formation and growth of NPs within glass matrices. The size-controlled preparation of NPs within glass matrices, which remains a challenge, is essential for advanced applications. Therefore, a thorough understanding of GNC synthesis techniques is expected to facilitate the preparation of innovative GNCs. KEYWORDS: exchange glass nanocomposites; melt-quenching; sol-gel; ion implantation; ion- Contents 1 2 Introduction Applications 2.1 Optoelectronics 2.1.1 Non-linear optical phenomena 2.1.2 Non-linear optical materials 2.1.2.1 Quantum dots embedded in glass matrices 2.1.2.2 Metal nanoparticles embedded in glass matrices 2.2 Photonics 2.3 Sensing Received March 5, 2022; accepted April 24, 2022 E-mail: fonsecagarcia.j@northeastern.edu 3 2.4 Electrochemistry 2.5 Catalysis 2.6 Biomedicine 2.7 Art Synthesis methods of GNCs 3.1 Melt-quenching 3.1.1 Introduction 3.1.2 Concomitant synthesis of glass matrix and nanoparticles 3.1.3 Precipitation of nanoparticles 3.1.3.1 Heat-assisted nanoparticle precipitation 3.1.3.2 Irradiation-assisted nanoparticle precipitation 3.1.4 Nanoparticles added to glass precursors before melting process 2 Front. Mater. Sci. 2022, 16(3): 220607 3.2 Sol-gel technique 3.2.1 Introduction 3.2.2 Concomitant synthesis of glass matrix and nanoparticles 3.2.3 Sol-gel technique combined with nanoparticle or nanoparticle precursors impregnation 3.3 Ion implantation 3.3.1 Introduction 3.3.2 Single implantation 3.3.3 Multiple implantation 3.4 Ion-exchange 3.4.1 Introduction 3.4.2 Direct formation of NPs 3.4.3 Thermal annealing in air 3.4.4 Thermal annealing in H2 atmosphere 3.4.5 Laser irradiation techniques 3.5 Less conventional techniques 3.5.1 Staining process 3.5.2 Spark plasma sintering 3.5.3 Radio frequency sputtering 3.5.4 Spray pyrolysis 3.5.5 Chemical vapor deposition 4 Conclusions Acknowledgements References 1 Introduction The concept of nanotechnology was introduced by Nobel laureate Richard P. Feynman at a meeting of the American Physical Society in 1959. In his famous lecture entitled “There’s Plenty of Room at the Bottom”, Feynman revealed his idea of manipulating matter at atomic scale. Since then, nanoscience and nanotechnology have grown rapidly and have been applied in almost every field related to industry [1‒3]. Nanoscience is defined as the study of matter at nanometer scales ranging from 1 to 100 nm. Nanotechnology is that technology that utilizes nanoscience in practical applications [4]. Subnanoclusters and nanoparticles (NPs) have attracted interest due to their remarkable properties compared to their bulk analogues. The changes in the properties of the materials in the nano- and subnano-regimes are related to their high surface-to-volume ratios and their morphology. However, subnanoclusters and NPs are often very unstable and tend to aggregate. The incorporation or embedding of subnanoclusters or NPs in host materials solves this problem [5‒8]. Logically, the properties of embedded NPs depend on the host matrix and the hostguest interactions [9]. Similarly, the geometric configurations and, in turn, the properties of the subnanoclusters are strongly affected by the host matrix in which they are incorporated. It should be mentioned that currently the incorporation of subnanoclusters with a narrow size distribution and a constant geometric structure in porous supports is a great challenge [10]. Glass is an antique invention of mankind. The oldest manufactured opaque glass, which was found in Egyptian potteries, dates from the eighth millennium BC [11]. The concept of glass has gradually evolved over time due to the progress of science and technology. Nowadays, glasses may be defined as nonequilibrium, non-crystalline condensed state of matter that spontaneously relaxes toward the supercooled liquid state. The structure of glasses is similar to that of their parent supercooled liquids. They crystallize in the limit of infinite time [12]. Glasses have been traditionally classified as inorganic, organic, and metallic, which exhibit ionic-covalent, covalent, and metallic bonds, respectively. Recently, a fourth family of glasses based on metal-organic frameworks (MOFs), which contain coordination bonds, has been reported [13]. Glasses exhibit high optical transparency, structural rigidity, compositional flexibility, property tailoring suitability, and durability. They have many applications ranging from traditional, such as windows, mirrors, containers, lighting, lenses, and handcrafted art, to technologically advanced, such as laser glass, radiation shielding glass, optical communication fibers, bioglass, armor glass, solar glass, etc. [11]. Glass composites are hybrid materials made of a glass matrix, which is the continuous phase, and the reinforcement(s) or dispersed phase(s), which is (are) present to a lesser extent. Glasses are excellent host materials. The glass matrices protect the dispersed phase(s) from the environment. The reinforcement may be either amorphous or crystalline. The purpose of the reinforcement(s) is to improve the properties of the glass matrix. Thus, the physical and chemical properties in the glass composites are different from those of the constituent phases [11]. Glass nanocomposites (GNCs) are defined as composite materials in which the matrix is a glass and at least one of the reinforcements is nanometric in size. Glasses are an exceptional matrix for NP growth due to their Javier FONSECA. NPs embedded into glass matrices: GNCs advantageous characteristics such as easy and inexpensive processing, high mechanical strength, high optical transparency, and excellent chemical durability [14‒15]. In fact, glasses with dispersed NPs were developed during the Roman Era [16]. The Lycurgus Cup, dating from the fourth century AD, is one of the most famous specimens. It is an early example of nanotechnology in human history. The Lycurgus Cup was made from silicate glass with Au and Ag NPs (50‒70 nm) incorporated within the matrix [17]. The properties of the GNCs depend on the glass matrix material and the reinforcement(s) size, shape, orientation, degree of loading, and dispersion. In GNCs, there is a strong reinforcement-interface interaction due to the high surface free energy of the dispersed phase. The remarkable properties of NPs embedded into glass matrices are associated with the small size effect, the large grain boundary effect, and the quantum confine effect. Thus, NPs may improve the mechanical properties (such as modulus, strength, and dimensional stability), thermal properties (such as conductivity, stability, flame retardancy, and heat distortion temperature), electrical properties (such as conductivity and capacitance), optical properties (such as absorption, scattering, and photoluminescence (PL)), chemical properties (such as reactivity, acid-alkali resistance, and durability) and surface properties (such as appearance) of glass matrices [11]. GNCs have attracted great interest driven by their large number of potential applications in fields such as optoelectronics, photonics, sensing, electrochemistry, catalysis, biomedicine, and art [18‒20]. The size, shape, number, and size distribution of the NPs determine the performance of the GNCs in all these fields [14]. Specifically, advanced applications require a well-defined particle size and size distribution. Size-controlled preparation of NPs within glass matrices remains a challenge. Understanding the kinetics growth of NPs is an essential prerequisite for improving NP syntheses in glass matrices. In this review, we focus on GNC synthesis protocols paying particular attention to the kinetics growth of NPs within glass matrices for each synthesis approach. Far from being obsolete materials, GNCs have an important impact on daily life, both at a purely commercial level and at the frontier research level. The ancient synergy between NPs and glasses can lead to the realization of intriguing new materials with interesting applications. We anticipate a significant increase in research efforts in the field of GNCs. We expect this 3 review serves as useful reading to motivate researchers at all levels to move or return to the GNC topic. In this review, we begin by describing the potential applications of GNCs in the fields of optoelectronics, photonics, sensors, electrochemistry, catalysis, biomedicine, and art. The fundamentals behind those fields are briefly explained to understand the suitability of GNCs in each application. In the next section of this Review, we extensively describe the most important synthesis methods of GNCs: melt-quenching, sol-gel, ion implantation, ionexchange, staining process, spark plasma sintering, radio frequency sputtering, spray pyrolysis, and chemical vapor deposition techniques. The aim of this extensive review is to explore the mechanisms of formation of GNCs. Many studies are reported for each GNC synthesis technique paying special attention to the underlying mechanism of GNC formation and the kinetics growth of NPs within glass matrices. We conclude this review by highlighting the importance of GNCs in various fields. 2 Applications In this section, we describe the potential applications of GNCs in the fields of optoelectronics, photonics, sensing, electrochemistry, catalysis, biomedicine, and art. By doing this, we hope to show the reader the tremendous possibilities of GNCs. 2.1 Optoelectronics The study of electronic devices that generate, detect and control light is known as optoelectronics (or optronics). Non-linear optical (NLO) glasses have attracted great attention for decades in the field of optoelectronics due to their high transparency, ease of manufacture, durability, and stability [21‒29]. The non-linearity in glasses is generally quite small, but the non-linear response can be improved by several orders of magnitude by embedding NPs. Particularly, all-optical switching devices based on NLO glasses with embedded NPs can potentially overcome the speed limitation imposed by electrical switching devices [30]. It is worth noting that optical devices can operate in a time range inaccessible to electronics (< 1 ps) [31]. In non-linear switching devices, the intensity of an optical signal changes between two output channels that perform logic operations (Fig. 1). All-optical switching devices include non-linear 4 Front. Mater. Sci. 2022, 16(3): 220607 [34‒35]. Polarization (P, dipole moment per unit volume) induced by an applied electric field (E) can be expressed as a Taylor expansion as follows: ( ) P = ε χ(1) E + χ(2) E 2 + χ(3) E 3 + ... Fig. 1 Examples of all-optical devices: (a) directional coupler (self-induced switching); (b) Mach‒Zehnder interferometer. The response of linear (dashed lines) and nonlinear (solid lines) materials is also illustrated. directional couplers, non-linear mode sorters, non-linear distributed couplers, non-linear Mach‒Zehnder interferometer, distributed feedback grating, etc. [32‒33]. Third-order NLO susceptibility (χ(3)), also known as optical Kerr susceptibility, assesses the applicability of any material in all-optical switching. Both the non-linear absorption coefficient (β ∝ Im[χ(3)]) and the non-linear refractive index (n2 ∝ Re[χ(3)]), which play a fundamental role in all-optical devices, are related to χ(3). The incorporation of reinforcements in glass matrices is currently being studied to improve the non-linearity of the pristine glass. Both quantum dots (QDs) and metal NPs embedded in glass matrices have been found to enhance χ(3). Specifically, metal NPs improve the χ(3) of the pristine glass at wavelengths very close to that of the characteristic localized surface plasmon resonance (LSPR) of the metal NPs [34‒35]. 2.1.1 Non-linear optical phenomena Non-linear susceptibility arises from the response of an electron to an electromagnetic field in an anharmonic potential well. The electromagnetic field of a light that propagates through an optical medium exerts a polarizing force on all the electrons that make up such medium. Logically, the polarization is exerted mainly on the external electrons. The polarization response is linear for weak electric fields of light waves (weaker than the intraatomic field which is 10 V·nm−1). However, the response becomes increasingly non-linear for electric fields comparable to or greater than the intra-atomic field (1) where ε is the dielectric constant of the medium, χ(1) is the linear dielectric susceptibility of the medium and χ(2), χ(3) and so on are the non-linear coefficients. χ(1) involves all the phenomena associated with linear optics, such as reflection, refraction, and interference. Generally, χ(1) is greater than χ(2) which, in turn, is greater than χ(3) and so forth [34‒35]. NLO phenomena can be classified as intrinsic (or structural) or extrinsic (or compositional). Intrinsic NLO phenomena include all light-induced changes in structure, such as changes in electronic density, average interatomic distances, molecular orientation, phase transition, etc. On the other hand, extrinsic NLO phenomena comprise all changes in chemical composition induced by light such as molecular dissociation, polymerization, etc. [34‒35]. As previously mentioned, n2 and β play a critical role in alloptical devices. In amorphous materials, e.g., glasses, n2 and β are related to the real and imaginary parts of χ(3), respectively [36]: 12 π [ (3) ] Re χ n0 (2) 96 π2 ω [ (3) ] Im χ n20 c2 (3) n2 = β= where Re[χ(3)] and Im[χ(3)] are the real and imaginary parts of χ(3), respectively; ω is the frequency of the applied electric field, c is the speed of light, n0 is the linear refractive index and is given by Eq. (4), and n2 and χ(3) are in e.s.u. ( [ ])0.5 n0 = 1 + 4 πRe χ(1) 2.1.2 (4) Non-linear optical materials A NLO medium must meet several basic requirements. Proper selection of a NLO material for practical application requires defining the application, device characteristics and operating conditions. Anyhow, materials with inherent χ(3) or high-intensity optical Javier FONSECA. NPs embedded into glass matrices: GNCs 5 field-induced χ(3) are always needed, regardless of practical application [34‒35]. The inherent χ(3) of a glass matrix can be improved by several orders of magnitude by embedding semiconductor or metal NPs into the glass [37]. These NPs have sizes smaller than the wavelength of light [38]. Electrons in these NPs are confined between infinite potential barriers, leading to the decomposition of the conduction and valence bands into a set of discrete levels [39]. This quantum confinement modifies the interaction of electrons with applied optical fields [40]. Furthermore, when the size of the NPs is much smaller (< λ/20) than the wavelength (λ) of the applied optical field, the electric field that polarizes the electrons of the NPs can be very different from the macroscopic field outside the NPs in the surrounding medium [41]. This effect is known as dielectric or classical confinement. Local field effects arise from dielectric confinement due to the difference in dielectric constants between the NPs and the surrounding glass matrix. In brief, the χ(3) of GNCs is affected by both quantum and dielectric confinement effects. 2.1.2.2 2.1.2.1 Quantum dots embedded in glass matrices The GNC non-linear susceptibility (χ(3) eff ) is related to the non-linear susceptibility of the metal NPs (χ(3) m ) according to the effective medium theory [45]: The effective dielectric constant ( ε) of a GNC (made of QDs embedded in a glass matrix), derived from effective medium theory in the self-consistent field approximation, can be expressed as [42]: (1 − p) εQD − ε εd − ε +p =0 2εQD − εd 2ε + εQD (5) where p is the volume fraction of the NPs, and εQD and εd are the dielectric constants of the embedded QDs and the glass matrix, respectively. The optical fields inside (Ein) and outside (Eex) the QDs are related: Ein = fQD · Eex (6) where fQD is a local field enhancement factor [43]. It can be expressed as: [ ( )]−1 εQD fQD = 1 + A −1 εd (7) where, A is a geometry-dependent factor: A = 1/3 for spherical particles; 1 > A > 1/3 for oblate spheroidal shapes; and 0 < A < 1/3 for prolate spheroidal shapes. Metal nanoparticles embedded in glass matrices ε of a GNC, made of metal NPs embedded in a glass matrix, is given by [44]: ε − εd εm − εd =p ε − 2εd εm + 2εd (8) where εm is the dielectric constant of the metal NP. Considerations regarding Eq. (8): (a) spherical NPs; (b) the size of the NPs is much smaller than the wavelength of light; (c) low-volume fraction (p << 1); and (d) absence of interactions among NPs. The optical fields inside (Ein) and outside (Eex) the metal NPs are related: Ein = fm · Eex (9) where fm is the local field enhancement factor. It describes the first-order susceptibility near or at LSPR of the metal NPs. The fm can be expressed as: fm = (3) 3εd εm + 2εd χeff = p · fm4 · χ(3) m (10) (11) where, χ(3) is proportional to the fourth power of the local eff field enhancement factor, determined at a frequency close to LSPR of metal NPs. Therefore, χ(3) eff increases considerably under LSPR conditions [46‒51]. The χ(3) m of the metal NPs includes contributions from (i) intra-band, (ii) interband, and (iii) hot-electron transitions: (i) Conduction electrons exhibit non-linear behavior in metal NPs due to the quantum-size effect [52]. This contribution vanishes in large particles. In other words, this process is highly dependent on size. (ii) Interband transitions between the d-levels and the conduction band also contribute to the χ(3) m . This process is resonant and does not depend on the size of NP [53]. (iii) Hot-electron transition. Strong absorption near the LSPR, which is dependent on particle size, can generate hot photo-excited electrons that change the Fermi‒Dirac distribution and thus the interband transition [54]. As explained above, the NLO response of GNCs with metal NPs is highly dependent on the LSPRs. The surface plasmon resonance (SPR) is the resonant oscillation of 6 Front. Mater. Sci. 2022, 16(3): 220607 conduction electrons at the interface between a negative dielectric material (metal) and a positive dielectric material (glass) stimulated by incident light. Specifically, LSPR is the collective oscillation of conduction electrons of metal NPs in resonance with the electromagnetic wave, which occurs when the metal NPs are smaller than the mean free path of electrons (Fig. 2) [55]. This localized field decays rapidly from the metal NPs-glass interface to the glass background. The LSPR frequency depends on the nature, size, and shape of the metal NPs [56‒61], the dielectric properties of the glass matrix [62‒63], the interNP coupling interactions [64‒66], and the wavelength of the incident light [67]. It is worth emphasizing that the electron cloud of the glass matrix often perturbates the electrons of the metal NPs [68‒69]. The resonance oscillations of the electron cloud of metal NPs allow them to be used as sensors. In brief, an analyte adsorbed on the surface of metal NPs generates a change in the electric field that can be detected as SPR shifts. Furthermore, rareearth (RE) ions experience local field enhancement when close to plasmonic NPs. This enhancement results in an increase in the emission intensities of RE ions. The improvement of luminescence by plasmonic NPs depends on the LSPR frequency [70‒72], and the orientation and distance of the luminophore from the metal NP [73‒74]. In addition, energy transfer between plasmonic NPs and RE ions also enhances the luminescence of those ions [75]. 2.2 Photonics Glasses doped with RE ions have been extensively investigated over the past decades due to their applicability in the field of photonics [76‒78]. Glasses are considered ideal hosts for RE ions due to their mechanical and thermal stability and cost-effective preparation. The PL properties of the RE ions within the glasses are limited by: (1) concentration quenching due to the solubility of Fig. 2 LSPR generated by incising light on a metal NP. the RE ions; and by (2) multiphonon relaxation due to the phonon energy of such glasses [79]. Therefore, glasses that allow a large amount of RE ions to be incorporated and exhibit low phonon energy are very desirable [80‒82]. In addition, the small β and low quantum efficiency of RE-doped glasses also restrict their applicability [83]. Metal NPs embedded within RE-doped glasses can overcome these difficulties, favoring the PL of RE ions and thus leading to high-performance photonic devices [84‒85]. Specifically, metal NPs, through LSPR, amplify the local electric field around the RE ions, enhancing the emission intensity [86‒88]. Therefore, GNCs doped with RE ions are promising photonic materials [89‒91]. More research in this field can be expected in the coming years. 2.3 Sensing Surface-enhanced Raman spectroscopy (SERS) is a promising analytical technique that enables sensitive detection with chemical specificity of a single molecule [92‒93]. SERS overcomes the low Raman scattering cross sections (10−30‒10−25) of the Raman effect [94]. It is worth noting that, in Raman experiments, high concentrations and high laser intensities are generally required to detect Raman scattering signals. SERS is based on LSPR at the interface of the surface of metal NPs and dielectric hosts [95‒97]. The LSPR generally enhanced the Raman modes of a molecule by a factor of η2, where η is the enhancement factor. Moreover, the emitted Raman signals are further improved by the same LSPR effect. Therefore, the output Raman signal is enhanced by a factor of η4 [98]. The maximum enhancement occurs near the plasmon frequency [99]. Non-spherical metal NPs such as metallic nanorods, nanowires, or nanofiber show improved SERS activity compared to spherical NPs [93,100]. In other words, the LSPR is further promoted by lightening rod effects (LREs) on the surface of non-spherical NPs. Since the discovery of SERS by Fleischman et al. in 1974 [101], great efforts have been made to develop inexpensive SERS-active substrates with high sensitivity, reproducibility, and stability [102‒105]. It is especially difficult to produce long-term stable SERS substrates with a high degree of reproducibility. In particular, GNCs are considered promising SERS-active substrates [106‒109]. Research on SERS-active GNCs is expected to be expanded in the near future. Javier FONSECA. NPs embedded into glass matrices: GNCs 2.4 Electrochemistry An immediate challenge for modern society is to reduce greenhouse gases by replacing vehicles that run on internal combustion engine with those that run on safe, low-cost rechargeable batteries with a high volumetric energy density and long life cycle [110]. Traditional rechargeable batteries use an aqueous electrolyte that rapidly transports H+ cations. The energy-gap “window” (Eg = 1.23 eV) of an aqueous electrolyte restricts the discharge voltage of a cell with a long shelf life to Vdis ≤ 1.5 V [111]. The organic-liquid Li+ electrolytes of the lithium-ion battery are inexpensive, easy to prepare, and have the required high conductivity of Li+. However, they are flammable. The formation and growth of dendrites across a liquid electrolyte to the cathode during charge can cause an internal short-circuit with incendiary consequences unless the dendrites are blocked [112]. The organic-liquid electrolyte has an energy-gap window of around 3 eV that limits the voltage of a cell with a long life cycle to Vdis ≤ 3.1 V [111]. Furthermore, if the battery cells have an anode with a chemical potential (Fermi level) lower than that of lithium metal, a Li+conductive solid-electrolyte interphase (SEI) passivating layer forms on the anode that prevents electrolyte reduction. The formation/dissolution of the SEI layer in a charge/discharge cycle restricts the life cycle of a cell [113]. The solid-state batteries (SSBs) that use solid electrolytes (SEs) can be an alternative [114]. Basically, a SSB is an electrochemical chain characterized by the continuity of cation transport from one electrode to another [115]. SEs offer some unique beneficial features: (1) Blocking dendrite growth and even preventing dendrite formation, thus eliminating the safety issue. (2) High current densities and fast charging times. (3) Stability at elevated temperatures. Thus, SEs improves battery safety. (4) No bulk polarization limiting the cell current. Li+ cation is the only mobile species. The associated anion is immobilized in a rigid macromolecular network. Therefore, there is no bulk polarization. SEs can be divided into inorganic solids (crystalline, glass or glass-ceramic in nature) and organic solid polymers [116]. However, the lithium-ion conductivity of the polymer electrolytes is too low for battery operation at room temperature. In addition, the rate capability of these electrolytes is limited, preventing fast charging [117]. Polymer-based batteries will certainly not enter the 7 market. On the other hand, the potential of glass SEs has been demonstrated [113]. Glass SEs are of special interest because various glass compositions exhibit good lithium conductivities (10−4 to 10−3 S·cm−1) at room temperature [118]. The critical question is whether alkali-metal anodes can be plated through a glass electrolyte efficiently at high rates for thousands of charge/discharge cycles. Logically, a low cost per kWh of energy storage capacity, which is highly dependent on manufacturing, is also a key requirement [119]. In brief, GNCs are attracting a lot of research attention recently due to their promising ability as SEs in SSBs [120‒127]. 2.5 Catalysis Catalysis is an essential technology to accelerate and direct chemical transformations [128‒129]. Heterogeneous catalysts have several advantages over homogeneous catalysts: easy catalyst separation, reuse, minimal product contamination, and high stability [130‒131]. Heterogeneous catalytic processes constitute 90% of all chemical production [132]. Specifically, supported metal NPs are the most common type of heterogeneous catalyst in industry [132]. Metal NPs are excellent catalysts in many industrially important chemical reactions [133‒134]. The size, shape, and surface functionalities of metal NPs control the activity of the metal NP-based catalysts [128,135]. Metal NPs, which exhibit extremely high surface energy, tend to aggregate and, thus, reduce their catalytic performance. Therefore, one of the critical functions of the support is to stabilize and immobilize the metal NPs. The support can also affect the catalytic efficiency of the metal NP and alter the reaction mechanism [128]. Glasses, which are inexpensive, environmentally friendly, durable, reusable, and inert, are candidates to replace conventional catalyst supports. The main drawback of glasses to be used as metal NP supports is their lack of porosity. It is worth mentioning that highsurface-area porous supports allow the diffusion of reagents, products, unreacted precursor molecules and byproducts. Consequently, porous glasses are gaining attention as support materials of metal NP catalysts. In other words, GNCs are attracting interest as catalyst [136‒140]. 2.6 Biomedicine Since their development in the late 1960s, bioactive 8 Front. Mater. Sci. 2022, 16(3): 220607 glasses (BGs) have been applied in many biomedical practices such as bone regeneration, soft tissue repair, dental grafting, and antibacterial treatment [141‒145]. It is worth mentioning that Bioglass® 45S5 (Na2O‒CaO‒ SiO2‒P2O5), which was first reported by Hench et al. in 1971 [146], has revolutionized tissue engineering. During the last decade, bioactive glass nanoparticles (BGNs) have gained attention due to their small size, large specific surface area, and large surface to volume ratio [147]. Specifically, the porosity and surface area of BGNs favor their applicability as carriers of drugs and other biomolecules. BGNs exhibit obvious advantages over BGs in drug delivery [148]. The incorporation of NPs into BGs and BGNs has been found to improve their healing qualities, their bioadhesive strength, and their anticancer potential. The treatment of malignant bone tumors has attracted the attention of many research groups in the field of bone tissue engineering [149‒150]. As innovative anticancer agents, NPs with anticancer potential are embedded in BGN. These NPs are released by the BGN and taken up by bone cancer cells. Ideally, NPs for this application have high toxicity to bone cancer cells and very low toxicity to healthy cells, promoting rapid bone healing [151]. BGs with embedded iron oxide NPs have also been developed for bone regeneration and anticancer purposes. It should be mentioned that, when exposed to a magnetic field, magnetic NPs generate a local temperature rise, which favors hyperthermia therapy [152‒153]. NPs have also been embedded in BGs and BGNs for antibacterial purposes [154]. A composite prepared by incorporating Cu NPs into BGNs (Cu NPs@BGNs) and an eggshell membrane (ESM) has shown great efficiency in wound healing. Cu NPs@BGN significantly improved the surface hydrophilicity and surface hardness of ESM. Moreover, Cu NPs@BGN stimulated angiogenesis, inhibited the growth of bacteria, and improved healing quality and formation of continuous and uniform epidermis layers [155]. 2.7 Art GNCs have been used for decorative purposes since ancient times. The Romans used Au NPs to color glasses as early as the 4th century BC [156]. Silver stain decoration, which is based on incorporating a surface layer of Ag NPs within a glass, was developed in Mesopotamia in the 8th century AD. This technique evolved towards the luster decoration consisting of a thin glass film, composed of a heterogeneous distribution of Ag and Cu NPs, deposited on a ceramic substrate [157]. The luster decoration exhibits peculiar optical properties, producing brilliant reflections of different colors and iridescence. Silver stain and luster decorations were carried over to the West through Spain during the 13th century [158‒159]. These decorative techniques began to be applied to color window glasses during the 13th century and became widespread in the 14th, 15th and 16th centuries. Thus, the glass windows of cathedrals were colored with metal NPs in Europe during the Middle Ages. In Italy, luster decoration was used to produce the well-known polychrome luster Renaissance pottery of Deruta and Gubbio [157]. Historical GNCs may suffer surface degradation due to centuries of exposure to weathering. Nowadays, studies on the degradation of GNCs are promoted to facilitate preservation and restoration of the Historical Heritage [160‒162]. 3 Synthesis methods of GNCs The most important synthesis methods of GNCs (meltquenching, sol-gel, ion implantation, ion-exchange, staining process, spark plasma sintering, radio frequency sputtering, spray pyrolysis, and chemical vapor deposition techniques) are described in detail in this section. Table 1 compares the different GNC synthesis methods. The mechanisms of GNC formation and the kinetics growth of NPs within glass matrices are reviewed. 3.1 3.1.1 Melt-quenching Introduction Melt-quenching technique is a long-standing, wellestablished and simple method to produce GNCs (Fig. 3). In fact, most GNCs have been prepared by this technique. In this process, an appropriate mixture of raw materials is prepared. The mixture is generally milled in a wet medium, such as acetone, to obtain a uniform powder. Then, it is dried in the air to evaporate the wet medium and calcined at a temperature below the melting point to release out the gaseous substances (moisture + gas). The calcined homogeneous mixture of raw materials is held at a certain temperature for a certain duration depending on the melting point of its components. In doing so, a Javier FONSECA. NPs embedded into glass matrices: GNCs Table 1 9 Comparison of the different GNC synthesis methods Merits GNC synthesis methods Meltquenching Advantage Disadvantage Sol-gel Ion implantation Ion-exchange Staining process Spark plasma sintering Radio frequency sputtering Short processing times, low temperature processing, control of NP size, control of NP size distribution Long-standing, Inexpensive, Short Inexpensive, Long-standing, Short inexpensive, low processing control of NP inexpensive, processing simple, short temperature times, low location within does not times, processing processing temperature the glass, high damage glass, continuous times, processing, concentration co-embedding processing continuous control of NP of NPs within various NPs processing location within the glass, does the glass, high not damage concentration glass, coof NPs within embedding the glass various NPs High Complex, long Expensive, no Limited to Limited to Expensive, Expensive, temperature processing control of NP some species some species complex, high complex, no processing, no times, size of NPs of NPs temperature control of NP control of NP discontinuous distribution, processing location within size, no control processing, no may damage the glass of NP size control of NP glass, may distribution, no location within generate control of NP the glass impurities location within the glass Fig. 3 Spray pyrolysis CVD Inexpensive, Simple, short short processing processing times, times, continuous continuous processing processing High temperature processing, low production rate High temperature processing, high purity precursors Schematic illustration of GNC preparation using melt-quenching technique. uniform fusion of molten materials is achieved. The homogenization of the molten GNC can be achieved by stirring. The molten GNC is poured in a preheated mold. The GNCs are generally quenched in air or in liquid such as liquid nitrogen or mercury. The GNC can also be quenched in contact with a solid material with high thermal conductivity to achieve ultrafast cooling rates. The obtained GNC is annealed near the glass transition temperature (Tg) to remove residual thermal stresses generated during quenching. Finally, the as-prepared GNCs are usually sawed, grinded, and polished for subsequent characterization or application. The formation of NPs and their size distribution within the glass matrices can only be partially controlled with this method. Therefore, when GNCs are prepared by melt-quenching technique, their properties are only slightly tailorable [11,163]. Three melt-quenching approaches are distinguished for embedding NPs in a glass matrix: (1) concomitant synthesis of glass matrix and NPs; (2) precipitation of NPs; and (3) addition of NPs to the glass precursors before the melting process. Table 2 shows examples of GNCs prepared by melt quenching techniques [75,97,164‒189]. These methods are reviewed in the following sections paying particular attention to the synthesis protocol. 3.1.2 Concomitant synthesis of glass matrix and nanoparticles Glass matrix and NPs can be prepared concomitantly by melt-quenching. Controlling the size of NPs is the great challenge of this approach. The size of NPs is often highly dependent on melting temperature and time. It should be Oxyfluoride glass (36.10 wt.% SiO2, 19.61 wt.% Al2O3, 14.91 wt.% ZnF2, 21.14 wt.% SrF2, 5.70 wt.% B2O3, and 2.54 wt.% Na2O) Oxyfluoride glass (35.06 wt.% SiO2, 19.04 wt.% Al2O3, 14.48 wt.% ZnF2, 20.53 wt.% SrF2, 5.54 wt.% B2O3, 2.47 wt.% Na2O, and 2.89 wt.% YbF3) Tellurite glass (85.70 wt.% TeO2, 11.58 wt.% ZnO, and 2.72 wt.% Er2O3) 2.49 0.007 0.03 2.37 (7.5) 5‒15 10‒35 (5) 0.38 2.49 (12) Ag NPs 5.14 5.23 ‒ 0.001 0.001 1.10 0.11 0.24 0.23 0.5‒1 Loading of NPs/wt.% in excess (3.5) (< 5) (< 5) 3‒9 (40) (40) 30‒40 30‒40 (20) 5‒20 (4.8) Size of NPs (min‒max (mean))/nm Ag NPs Ag NPs Au NPs Au NPs Au NPs Au NPs Au NPs Cu NPs Sodium borosilicate glass (62.48 wt.% Ag NPs SiO2, 8.78 wt.% B2O3, 11.72 wt.% Na2O, 3.97 wt.% NaF, and 13.06 wt.% SrO) Sodium borosilicate glass (62.48 wt.% Ag NPs Ag NPs@SmF3-doped sodium borosilicate SiO2, 8.78 wt.% B2O3, 11.72 wt.% Na2O, glass 3.97 wt.% NaF, 13.06 wt.% SrO, and 2.49 wt.% SmF3) Bi-coated Ag Bismuthate glass (79.13 wt.% Bi2O3, Ag NPs coated NPs@bismuthate glass 12 wt.% K2O, and 8.87 wt.% B2O3) with Bi Bi-coated Ag Bismuthate glass (79.13 wt.% Bi2O3, Ag NPs coated NPs@bismuthate glass 12 wt.% K2O, and 8.87 wt.% B2O3) with Bi Ag NPs@calcium Calcium fluorophosphate glass Ag NPs fluorophosphate glass (12.63 wt.% CaO, 4.40 wt.% CaF2, 79.94 wt.% P2O5, and 3.03 wt.% SnO) Ag NPs@sodium borosilicate glass Ag NPs@tellurite glass Ag NPs@oxyfluoride glass Ag NPs@oxyfluoride glass Au NPs@bismuthate glass Au NPs@lanthanum borate glass Au NPs@lanthanum borate glass Au NPs@lanthanum borate glass Au NPs@lanthanum borate glass Cu NPs@tellurite glass Cu NPs@tellurite glass Lanthanum borate glass (23.05 wt.% La2O3, 47.38 wt.% PbO, and 29.56 wt.% B2O3) Lanthanum borate glass (23.05 wt.% La2O3, 47.38 wt.% PbO, and 29.56 wt.% B2O3) Lanthanum borate glass (23.05 wt.% La2O3, 47.38 wt.% PbO, and 29.56 wt.% B2O3) Lanthanum borate glass (23.05 wt.% La2O3, 47.38 wt.% PbO, and 29.56 wt.% B2O3) Bismuthate glass (91.22 wt.% Bi2O3, 6.82 wt.% B2O3, and 1.96 wt.% SiO2) CdSe QDs Silicate glass (56 wt.% SiO2, 8 wt.% B2O3, 24 wt.% K2O, 3 wt.% CaO, and 9 wt.% BaO) Tellurite glass (87.81 wt.% TeO2, 7.76 wt.% Sb2O3, 2.33 wt.% Yb2O3, 0.97 wt.% Ce2O3, and 1.13 wt.% Er2O3) Tellurite glass (90.94 wt.% TeO2, 7.91 wt.% Sb2O3, and 1.15 wt.% Er2O3) Cu NPs Reinforcement Host matrix Composite CdSe QDs@silicate glass GNCs prepared by melt-quenching techniques [75,97,164–189] Table 2 ‒ Applications Heat-assisted precipitation of NPs: meltquenching for glass preparation (melting at 1200 °C for 15 min); heat-assisted NP precipitation (at 550 °C for 10 h) Concomitant synthesis of glass matrix and NPs: melt-quenching (melting at 1100 °C for 10 min); annealing (at 350 °C for 2 h) Concomitant synthesis of glass matrix and NPs: melt-quenching (melting at 800 °C for 15 min); annealing (At 350 °C for 8 h) Concomitant synthesis of glass matrix and NPs: melt-quenching (melting at 1450 °C for 1 h); annealing (at 450 °C for 2 h) Concomitant synthesis of glass matrix and NPs: melt-quenching (melting at 1200 °C for 30 min); annealing (at 390 °C for 2 h) Concomitant synthesis of glass matrix and NPs: melt-quenching (melting at 1450 °C for 45 min) Concomitant synthesis of glass matrix and NPs: melt-quenching (melting at 1100 °C for 30 min); annealing (at 500 °C for 2 h) ‒ Nanophotonics and optoelectronics Solid-state lasers Nanophotonics LED illumination and solar cells NLO devices: χ(3)(800 nm) = 4.88 × 10−10 esu NLO devices: n2(532 nm) = 5.6 × 10−14 m2·W−1 NLO devices: n2(532 nm) = 6.7 × 10−14 m2·W−1 Concomitant synthesis of glass matrix and NPs: IR optical devices melt-quenching (melting at 800 °C for 3 min), Sb2O3 redox route Concomitant synthesis of glass matrix and NPs: Lasers (biomedical and melt-quenching (melting at 800 °C for 3 min), bioimaging fields); Sb2O3 redox route thermophotovoltaic conversion of thermal radiation Concomitant synthesis of glass matrix and NPs: NLO devices: β(532 nm) = melt-quenching (melting at 1100 °C for 2 × 10−11 m·W−1 30 min); annealing (at 500 °C for 2 h) NLO devices: β(532 nm) = 1.4 × 10−12 m·W−1 Concomitant synthesis of glass matrix and NPs: melt-quenching (melting at 1400 °C for 1.5 h) Melt-quenching technique [173] [172] [171] [97] [170] [169] [168] [167] [166] [165] [164] Ref. 10 Front. Mater. Sci. 2022, 16(3): 220607 Au NPs Silicate glass (70.02 wt.% SiO2, 9.34 wt.% CaO, and 20.64 wt.% Na2O) Silicate glass (70.02 wt.% SiO2, Ag NPs and Au 9.34 wt.% CaO, and 20.64 wt.% Na2O) NPs Ag NPs Silicate glass (70.02 wt.% SiO2, 9.34 wt.% CaO, and 20.64 wt.% Na2O) Borate glass (6.78 wt.% K2O, 4.04 wt.% CaO, and 89.19 wt.% B2O3) Soda-lime silicate glass (46.62 wt.% SiO2, 18.37 wt.% B2O3, 7.19 wt.% MgO, 9.62 wt.% Na2O, and 18.20 wt.% Al2O3) Borosilicate glass (75.22 wt.% SiO2, 16.98 wt.% Na2O, 3.01 wt.% Al2O3, and 4.79 wt.% CaO) Borosilicate glass (72.68 wt.% SiO2, 15.67 wt.% Na2O, 2.19 wt.% Al2O3, and 9.45 wt.% CaO) Ag NPs@silicate glass Au NPs@silicate glass Ag NPs/Au NPs@silicate glass Cu NPs@borate glass Cu NPs@soda-lime silicate glass Ag NPs@borosilicate glass Au NPs@borosilicate glass (9) Cu NPs (1.8) 1‒5.6 (2.8) 3‒5 (4) ‒ 6‒8 1‒8 Cu NPs Au NPs Ag NPs Ag NPs and Au NPs Phosphate glass (82.58 wt.% P2O5, 3.69 wt.% MgO, 11.10 wt.% ZnSO4, and 2.63 wt.% Eu2O3) Ag NPs/Au NPs@phosphate glass (27.65) (6.40) Ag NPs@lanthanum sodium borate glass Ag NPs Lanthanum sodium borate glass (67.23 wt.% B2O3, 11.14 wt.% Na2O, 19.52 wt.% La2O3, and 2.11 wt.% Eu2O3) Ag NPs@lanthanum sodium borate glass (2.72) Ag NPs Lanthanum sodium borate glass (67.23 wt.% B2O3, 11.14 wt.% Na2O, 19.52 wt.% La2O3, and 2.11 wt.% Eu2O3) Lanthanum sodium borate glass (67.23 wt.% B2O3, 11.14 wt.% Na2O, 19.52 wt.% La2O3, and 2.11 wt.% Eu2O3) Ag NPs@lanthanum sodium borate glass (42) Ag NPs Size of NPs (min‒max (mean))/nm (15) (4.64) Calcium fluorophosphate glass (12.63 wt.% CaO, 4.40 wt.% CaF2, 79.94 wt.% P2O5, and 3.03 wt.% SnO) Ag NPs@calcium fluorophosphate glass Ag NPs Reinforcement Ag NPs Calcium fluorophosphate glass (12.63 wt.% CaO, 4.40 wt.% CaF2, 79.94 wt.% P2O5, and 3.03 wt.% SnO) Host matrix Ag NPs@calcium fluorophosphate glass Composite Melt-quenching technique Heat-assisted precipitation of NPs: meltquenching for glass preparation (melting at 1200 °C for 15 min); heat-assisted NP precipitation (at 550 °C for 30 h) 2.37 Heat-assisted precipitation of NPs: meltquenching for glass preparation (melting at 1200 °C for 15 min); heat-assisted NP precipitation (at 550 °C for 50 h) 0.17 Heat-assisted precipitation of NPs: meltquenching for glass preparation (melting at 1050 °C for 2 h); heat-assisted NP precipitation (at 450 °C for 10 h) 0.17 Heat-assisted precipitation of NPs: meltquenching for glass preparation (melting at 1050 °C for 2 h); heat-assisted NP precipitation (at 450 °C for 15 h) 0.17 Heat-assisted precipitation of NPs: meltquenching for glass preparation (melting at 1050 °C for 2 h); heat-assisted NP precipitation (at 450 °C for 20 h) Heat-assisted precipitation of NPs: melt0.44 wt.% Ag, quenching for glass preparation (melting at 1.5 wt.% Au 1100 °C for 1.5 h); heat-assisted NP precipitation (at 300 °C for 3 h) Irradiation-assisted precipitation of NPs: melt‒ quenching for glass preparation (melting at 1550 °C for 1 h); femtosecond laser irradiation; annealing (at 550 °C for 10 min) Irradiation-assisted precipitation of NPs: melt‒ quenching for glass preparation (melting at 1550 °C for 1 h); femtosecond laser irradiation; annealing (at 550 °C for 30 min) Irradiation-assisted precipitation of NPs: melt‒ quenching for glass preparation (melting at 1550 °C for 1 h); femtosecond laser irradiation; annealing Irradiation-assisted precipitation of NPs: melt‒ quenching for glass preparation (melting at 1000 °C for 30 min); annealing (at 300 °C for 2 h); femtosecond laser irradiation Irradiation-assisted precipitation of NPs: melt‒ quenching for glass preparation (melting at 1400 °C for 1 h); annealing (at 400 °C for 12 h); femtosecond laser irradiation; heat treatment (at 600 °C for 1 h) Irradiation-assisted precipitation of NPs: melt‒ quenching for glass preparation (melting at 1450 °C for 4 h); X-ray irradiation (34 min); heat treatment (at 400 °C for 15 min) Irradiation-assisted precipitation of NPs: melt‒ quenching for glass preparation (melting at 1400 °C for 3 h); annealing (at 550 °C for 15 min); X-ray irradiation (5 min); heat treatment (at 600 °C for 20 min or for 3 h) Loading of NPs/wt.% in excess 2.37 Art, optical memory, and alloptical switching devices Art, optical memory, and alloptical switching devices Art, optical memory, and alloptical switching devices Art, optical memory, and alloptical switching devices Art, optical memory, and alloptical switching devices Art, optical memory, and alloptical switching devices Art, optical memory, and alloptical switching devices Solid-state lasers Solid-state lasers Applications [182] [181] [180] [179] [178] [177] [176] [175] [174] Ref. (continued) Javier FONSECA. NPs embedded into glass matrices: GNCs 11 Borate glass (58.10 wt.% B2O3, 31.12 wt.% ZnO, and 10.78 wt.% Na2O) Bio-borosilicate glass (10.17 wt.% SiO2, 14.15 wt.% B2O3, and 75.68 wt.% TeO2) Antimony phosphate glass (68.47 wt.% SbPO4, 16.47 wt.% ZnO, and 15.06 wt.% PbO) NiO NPs@borate glass Sm2O3 NPs@bioborosilicate glass CdFe2O4/SiO2 NPs@antimony phosphate glass Tellurite glass (86.97 wt.% TeO2, 10.17 wt.% ZnO, 0.21 wt.% LiO, and 2.66 wt.% Er2O3) Phosphate glass (77.24 wt.% P2O5, 19.68 wt.% ZnO, and 3.08 wt.% Er2O3) Phosphate glass (77.10 wt.% P2O5, 19.79 wt.% ZnO, and 3.10 wt.% Er2O3) Fe3O4 NPs@tellurite glass Fe3O4 NPs@phosphate glass Fe3O4 NPs@phosphate glass a) Size of NPs before being incorporated into the glass. Tellurite glass (88.44 wt.% TeO2, and 11.56 wt.% ZnO) Fe3O4 NPs@tellurite glass (26) Fe3O4 NPs 2.35 18‒70(a) 3.6222 2.72 3.18 18‒70(a) (31) 2.5 (3.5) 8.37 0.52 (70)a) (71.74) 0.4 ‒ Loading of NPs/wt.% in excess (4.93) 2‒8 Size of NPs (min‒max (mean))/nm Fe3O4 NPs Fe3O4 NPs Fe3O4 NPs CdFe2O4/SiO2 NPs Sm2O3 NPs NiO NPs Ag NPs Cu NPs Silicate glass (75.32 wt.% SiO2, 16.8 wt.% Na2O, 7.6 wt.% CaO, and 0.28 wt.% SnO2) Phosphate glass (77.46 wt.% P2O5, 3.79 wt.% MgO, 15.19 wt.% ZnSO4, and 3.56 wt.% Ho2O3) Cu NPs@silicate glass Ag NPs@phosphate glass Reinforcement Host matrix Composite Applications Irradiation-assisted precipitation of NPs: melt‒ quenching for glass preparation; ion irradiation; heat treatment (at 400 °C for 10 h) NPs added before glass formation: meltSolid-state lasers quenching for glass preparation (melting at 1100 °C for 1.5 h); annealing (at 300 °C for 3.5 h) NPs added before glass formation: meltSemiconductor applications quenching for glass preparation (melting at 1200 °C for 2 h); annealing (at 400 °C for 1 h) NPs added before glass formation: meltNon-linear optics quenching for glass preparation (melting at 1050 °C for 3 h); annealing (at 400 °C for 2 h) NPs added to a molten glass: melt-quenching Magneto-optical devices for glass preparation (melting at 1100 °C for 30 min); NPs added to a molten glass; meltquenching for embedding CdFe2O4/SiO2 NPs (melting at 1100 °C for 7 min) NPs added before glass formation: meltMagneto-optical devices quenching for glass preparation (melting at 850 °C for 20 min); annealing (at 300 °C for 3 h) NPs added before glass formation: meltMagneto-optical devices quenching for glass preparation (melting at 850 °C for 20 min); annealing (at 300 °C for 3 h) NPs added before glass formation: meltMagneto-optical devices quenching for glass preparation (melting at 950 °C for 1 h); annealing (at 300 °C for 3 h) Melt-quenching technique [189] [188] [187] [186] [185] [184] [75] [183] Ref. (continued) 12 Front. Mater. Sci. 2022, 16(3): 220607 Javier FONSECA. NPs embedded into glass matrices: GNCs noted that the high melting temperatures used to obtain the glasses can dissolve NPs. Franco et al. have developed a melt-quenching approach, referred to as the Sb2O3 redox route, to obtain Cu NPs embedded in a glassy matrix (NaPO3‒Sb2O3‒ CuO) [190]. The precursors were melted at 1000 °C under atmospheric conditions. Subsequently, the melt was quenched. Both Cu2+ and Cu+ ions, as well as Cu0, were present in the glass. The reducing properties of Sb2O3 during melting were used to decrease the oxidation number of copper ions in the glassy matrix. In other words, the oxidation reaction Sb3+ → Sb5+ + 2e− during the melting process induced the reduction of Cu2+ ions. Thus, the presence of Cu NPs was favored by high Sb2O3 concentration in the glasses. Moreover, the average particle size was found to increase with increasing reducing agent content. Therefore, control of the Cu NP size was possible through composition adjustments. The presence of Cu+ can be explained by an incomplete reduction process from Cu2+ to Cu0 species, leading to the formation of Cu2O, or by a partial oxidation process of Cu NPs, forming Cu‒Cu2O core‒shell NPs near to the surface of the glass. The GNCs with different contents of Sb2O3 were doped with Er2O3. Cu NPs enhanced the fluorescence of the Er3+ ions [190]. The Sb2O3 redox route has also been applied to incorporate Cu NPs into a tellurite glass (TeO2‒Sb2O3‒CuO) [165]. The glasses were prepared by melting the precursors at 800 °C for 3 or 20 min and then quenching. The formation of Cu NPs embedded in tellurite glasses was found to depend on the melting time. Cu NPs with dimensions ranging from 5 to 20 nm were observed when the melting time was 3 min. However, when the melting time increased, Cu NPs solubilized leading to the formation of Cu2+. The tellurite glass (obtained through 20 min of melting) and the Cu NPs@tellurite glass (prepared through 3 min of melting) were doped with Er3+, Yb3+ and Ce3+, which produced the composites called TeSbCuYCE and TeSbCuYCE + Cu NP, respectively. The influence of Cu NPs in the infrared emission Er3+: 4I13/2 → 4I15/2 at 1.55 µm in TeSbCuYCE + Cu NP was investigated. It is worth mentioning that the Yb3+ ion is an activator for this erbium transition and the Ce3+ ion increases the 1.55 µm transition lifetime and emission [191‒194]. The LSPR of the Cu NPs enhanced the 1.55 µm Er3+ emission by 47%. Moreover, the lifetime of the Er3+: 4I13/2 → 4I15/2 transition increased by about 50% in the presence of Cu NPs. TeSbCuYCE + Cu NP can be used in infrared optical devices [165]. 13 Er3+-doped TeO2‒Sb2O3‒CuO glass with embedded Cu NPs, which has also been prepared by the Sb2O3 redox route, has exhibited vibronic transitions of Er3+ ions at room temperature. This unprecedented behavior can be applied in biomedical and bioimaging fields and in thermophotovoltaic conversion of thermal radiation [166]. Rajaramakrishna et al. have embedded Au NPs in a lanthanum borate glass (La2O3‒PbO‒B2O3) by meltquenching technique [167‒168]. A mixture of raw materials was melted at 1100 °C for 30 min. The molted GNC was quenched at 200 °C and subsequently annealed at 500 °C for 2 h. All Au3+ ions were found to be reduced to Au0 after melt-quenching. The size of the Au NPs in the as-prepared GNCs ranged between 2 and 9 nm. These NPs grew at 40 nm when annealed. At 532 nm, β of the composites with 1.1 and 0.11 wt.% Au NPs was 1.4 × 10−12 and 2 × 10−11 m·W−1, respectively [167]. At 532 nm, n2 of the composites with 0.001 wt.% Au NPs annealed at 500 °C for 2 and 4 h was 6.7 × 10−14 and 5.6 × 10−14 m2·W−1, respectively [168]. Similarly, Au NPs have also been embedded into a bismuthate glass (Bi2O3‒B2O3‒SiO2) using melt-quenching approach [169]. Raw materials were melted at 1200 °C for 30 min and subsequently quenched. The quenched composite was annealed at 390 °C for 2 h. The size of the Au NPs in the glass ranged between 3 and 9 nm (Fig. 4). This GNC showed the χ(3) of 4.88 × 10−10 esu at 800 nm, which was of two orders of magnitude higher than that of the pristine glass [169]. Therefore, these GNCs (Au NPs@lanthanum borate glass and Au NPs@bismuthate glass) are potentially useful for NLO devices. Ag NPs have been formed within an Yb3+-doped oxyfluoride glass by a melt-quenching method [170]. The precursors were mixed, and then melted at 1450 °C for 45 min. The molted GNC was quenched at room temperature. Ag+ ions were reduced to Ag0 atoms during the melt-quenching process. Ag nanoclusters (in the Agmn+ form) smaller than 5 nm were suggested to be homogeneously distributed within the glass matrix. The GNC exhibited broad absorption and emission bands in the ultraviolet and visible spectral range, respectively. The GNC displayed a luminescence quantum yield (QY) of 96.75%. Due to the energy transfer from Ag NPs to Yb3+, the NIR emission of this GNC (Ag NPs embedded in an Yb3+-doped oxyfluoride glass) was four times stronger than that of Yb3+-doped oxyfluoride glass. This GNC is a promising candidate for LED illumination and solar cells [170]. In a similar study, Dousti et al. have embedded Ag 14 Front. Mater. Sci. 2022, 16(3): 220607 Fig. 4 (a) TEM image of the GNC (Au NPs@bismuthate glass). (b) HRTEM image of Au NP embedded into the glass matrix. Reproduced with permission from Ref. [169] (Copyright 2011 Elsevier). NPs in an Er3+-doped tellurite glass (TeO2‒ZnO‒Er2O3) using melt-quenching method (Fig. 5) [97]. The precursor mixture was melted at 800 °C for 15 min and subsequently quenched and annealed at 350 °C for 8 h. Ag NPs were formed and grown during melting and annealing, respectively. The average size of Ag NPs embedded in the glass matrix was approximately 12 nm. The Ag NPs showed a non-spherical shape. Ag NPs enhanced the Raman and PL intensities (of Er3+) eight and five times, respectively. It was suggested that high concentration of Ag NPs improves Raman and PL intensities in three ways: (1) providing the “hot-spots”, (2) increasing the coalescence of the NPs, and (3) favoring the energy transfer from NP to fluorophore. The authors also observed that quenching of PL emission in the visible range occurred after further annealing due to dissolution of aggregated NPs. In brief, improving the upconversion luminescence and Raman intensity in RE doped glasses contributes to the development of nanophotonics [97]. Guo et al. have investigated the effects of SmF3 doping on the formation and growth of Ag species (including Ag+ ions, Ag nanoclusters and Ag NPs) in a sodium borosilicate glass (SiO2‒B2O3‒Na2O‒NaF‒SrO) [171]. Ag species embedded in the SmF3-doped glass were prepared by melt-quenching. A mixture of raw materials was melted at 1450 °C for 1 h and subsequently quenched and annealed at 450 °C for 2 h. It should be noted that while the average size of Ag NPs in the undoped glass was around 3.5 nm, the mean size of Ag NPs in the SmF3doped glass increased to 7.5 nm. Sm3+ and Sm2+ ions were suggested to accelerate the formation and growth of Ag NPs as follows: Sm3+ + e− → Sm2+ (12) Fig. 5 (a) UV-vis spectra of GNCs. Notation: A05H8 = Composite made from Ag NPs (0.5 mol.% AgCl precursor) embedded into an Er3+-doped tellurite glass; A10H8 = Composite made from Ag NPs (1 mol.% AgCl precursor) embedded into an Er3+-doped tellurite glass. E1A05H8 revealed three prominent plasmon peaks at 562, 598, and 628 nm. E0A10H8 showed two peaks at 550 and 578 nm. This difference between E1A05H8 and E0A10H8 was attributed to the non-spherical Ag NPs. (b) Size distribution of Ag NPs (0.5 mol.% AgCl precursor) embedded into an Er3+doped tellurite glass. (c) TEM image of Ag NPs (0.5 mol.% AgCl precursor) embedded into an Er3+-doped tellurite glass. (d) HRTEM micrograph of Ag NPs (0.5 mol.% AgCl precursor) embedded into an Er3+-doped tellurite glass. The image shows a 0.125 nm fringe spacing corresponding to the (3 1 1) Ag lattice spacing. Reproduced with permission from Ref. [97] (Copyright 2013 Elsevier). Sm2+ + Ag+ → Sm3+ + Ag0 (13) Ag0 + Ag+ → Ag+2 (14) Ag0 + Ag+2 → Ag+3 (15) Similarly, the self-reduction of Eu3+ to Eu2+ in a glass matrix has also been reported to promote the formation and growth of Ag NPs within the glass [83,195]. Furthermore, the F− vacancies within the glass matrix were attractors of Ag+ ions, which benefited the formation of bigger Ag NPs. The annealing also favored the growth of Ag NPs in bigger ones. The luminescence properties of the composite were found to vary with the growth of Ag NPs [171]. Complex NPs have also been embedded within glass matrices, thus forming GNCs. Worsch et al. have prepared multicore magnetite NPs within a glass matrix (K2O‒Al2O3‒B2O3‒SiO2) by melt-quenching [196]. A Javier FONSECA. NPs embedded into glass matrices: GNCs mixture of raw materials was melted at 1400 °C for 2 h and subsequently at 1480 °C for 20 min. Finally, the molten glass was quenched. In doing so, GNC droplets with diameters between 100 and 150 nm were formed. Magnetite crystals ranging in size from 10 to 30 nm were prepared within the glass matrix. In this work, multicore magnetite NPs were extracted from the glassy matrix by boiling the pulverized composite in a concentrated aqueous solution of NaOH. The size of the multicore magnetite NPs was reported to range from 80 to 250 nm. The NPs exhibited superparamagnetic behavior. In addition, they showed a narrow hysteresis loop and a ratio of remanent versus saturation magnetization that was not high enough for uniaxial anisotropy. The temperaturedependent magneto-relaxometry (TMRX) measurements showed peaks at 13 and 39 K in the distribution of the relaxation of magnetic moments [196]. In brief, the crystallization of a glass containing Fe2O3−x allows the preparation of magnetite within the glass with a narrow size distribution and a large magnetic moment [197‒198]. Magnetite NPs can be subsequently extracted by dissolving or boiling the glass matrix [199]. Singh et al. have embedded Bi-coated Ag NPs in a bismuthate glass (K2O‒Bi2O3‒B2O3) using meltquenching technique [172]. The raw materials were mixed and subsequently melted at 1100 °C for 10 min. During the melting, the Bi3+ ions of Bi2O3 were thermally reduced as follows: Bi3+ → Bi2+ → Bi+ → Bi0 (16) In the presence of Bi2+ or Bi+, the Ag+ ions were reduced to Ag0: Bi2+ /Bi+ + Ag+ → Bi3+ /Bi2+ + Ag0 (17) Therefore, Ag+ ions were first reduced to Ag0 and then Bi3+/Bi2+/Bi+ ions were reduced to Bi0. In brief, Ag NPs were initially generated and later Bi0 coated the Ag NPs. Finally, the molten composite was quenched and then annealed at 350 °C for 2 h. Ag NPs were found to be spherical with a size ranging from 5 to 15 nm when the composite was doped with 0.007 wt.% Ag NPs. However, when the composite was doped with 0.03 wt.% Ag NPs, the Ag NPs were reported to be hexagonal with a size between 10 to 35 nm. The size of spherical Ag NPs increased with heat treatment at 360 °C. On the contrary, the size of the hexagonal Ag NPs remained constant after such treatment for up to 65 h. It is worth mentioning that 15 the GNCs were dichroic. The different orientations of NPs were suggested to be responsible for the dichroic nature of the composite. These GNCs have potential applications in nanophotonics and optoelectronics [172]. 3.1.3 Precipitation of nanoparticles After glass formation by melt-quenching, NPs can be precipitated within the as-prepared glass matrix by applying an external stimulus. Two principal precipitation techniques are distinguished: (1) heat-assisted NP precipitation; and (2) irradiation-assisted NP precipitation. 3.1.3.1 Heat-assisted nanoparticle precipitation Metal ions (Mn+) are present in a glass matrix at temperatures as high as those reached when it melts. Therefore, when a molten glass is quenched, the glass matrix becomes supersaturated with respect to Mn+. The heat treatment favors the diffusion and aggregation of Mn+ within the glass matrix, which leads to the formation of metal NPs. Besides metal NPs, other NPs such as QDs and nanocrystals (NCs) can also be precipitated within a glass matrix. In these cases, nuclei are generally produced during quenching, and as with metal NPs, the heat treatment facilitates coalescence and growth of these nuclei. Ag NPs have been embedded in calcium fluorophosphate glass (CaO‒CaF2‒P2O5) via melt-quenching technique and heat treatment [173]. A mixture of raw materials was melted at 1200 °C for 15 min and then quenched at room temperature. To precipitate Ag NPs, the quenched glasses were heated at 550 °C for 10, 20, 30, 40, and 50 h. The growth of Ag NPs in the glass matrix was found to be time dependent. Nucleation and growth theory explained the growth mechanism of Ag NPs within the glass matrix: • Ag+ and Sn2+ ions were present in the glass matrix at high temperature during melting. SnO was added to prevent nucleation and crystal growth of unwanted Ag2O [200]. • The glass system became supersaturated with respect to Ag+ when the molten glass was rapidly quenched. • Ag+ ions were reduced to Ag0 when the composite was heated at 550 °C. In other words, the quenched composites were heated to precipitate Ag NPs [201]. The average size of the NPs within a GNC heated for 10, 30, and 50 h was approximately 5, 15, and 42 nm, 16 Front. Mater. Sci. 2022, 16(3): 220607 respectively (Fig. 6). As the heating duration increased: (1) the GNC became compact; (2) Vickers microhardness and fracture toughness increased while brittleness decreased; (3) the full width at half maxima (FWHM) of the SPR in the visible region of the composite was broadened; and (4) higher dielectric constant and lower dielectric loss were observed for the GNC [173]. Ag NPs have also been embedded in Eu3+-doped lanthanum sodium borate glass (B2O3‒Na2O‒La2O3‒Eu2O3) by melt-quenching and heat-assisted precipitation [174]. A mixture of precursors was melted at 1050 °C for 2 h and subsequently quenched at 150 °C. The melt-quench material was annealed at 250 °C for 4 h. Finally, the glasses were heated at 450 °C for different times (5, 10, 15, and 20 h) allowing the nucleation and growth of Ag NPs. Ag NPs were found to be homogeneously distributed `within the glass matrix. The average size of the Ag NPs within the GNCs prepared with a heating time of 10, 15, and 20 h was approximately 2.72, 4.64, and 6.40 nm, respectively. Ag NPs enhanced the Eu3+ ion luminescence properties. Specifically, the optimum PL was obtained by the composite heated for 10 h. PL decreased in those Fig. 6 (a) TEM image of Ag NPs embedded in calcium fluorophosphate glass (heated at 550 °C for 10 h). Inset: the size distribution of Ag NPs. (b) HRTEM of Ag NPs@calcium fluorophosphate glass (heated at 550 °C for 10 h). (c) TEM image of Ag NPs embedded in calcium fluorophosphate glass (heated at 550 °C for 30 h). Inset: the size distribution of Ag NPs. (d) Selected area electron diffraction (SAED) image of Ag NPs embedded in calcium fluorophosphate glass (heated at 550 °C for 30 h). Reproduced with permission from Ref. [173] (Copyright 2015 Elsevier). composites heated more than 10 h. In other words, the PL quenching was more significant for the composites with longer heating time. The radiative parameters, which were calculated applying the Judd‒Ofelt theory, and the luminescence results suggest that the composite heated for 10 h is an appropriate material for the preparation of highefficiency LEDs and solid-state lasers [174]. In a similar study, Ag and Au NPs have been co-embedded in a Eu3+doped phosphate glass (P2O5‒MgO‒ZnSO4‒Eu2O3) by melt-quenching and precipitation [175]. Firstly, the precursors were melted at 1100 °C for 1.5 h. Then, the molten glass was quenched and finally heated at 300 °C for 3 h, which favored the precipitation of NPs. The average diameter of NPs within the glass matrix was reported to be 27.65 nm. The stability of the glass matrix decreased as the concentration of Au NPs increased. It was attributed to the breaking of many bridging oxygen (BO) linkages and the generation of more non-bridging oxygens (NBOs). The high quantum efficiency, the excellent CIE color purity, and the PL intensity for the red peak suggested that this composite is a great red laser prospect [175]. Zhang et al. have embedded Ag NPs and CsPbBr3 QDs in a sodium borosilicate glass by melt-quenching and precipitation approach [202]. The mixture of glass precursors was melted in an air atmosphere at 1200 °C for 15 min and subsequently quenched. To release thermal stresses, the glass was annealed at 350 °C for 10 h. Finally, the annealed glass was heated at 475 °C for 5 h to precipitate Ag NPs and CsPbBr3 QDs. Both Ag NPs and CsPbBr3 QDs coexisted in the glass matrix. They were reported to be almost uniformly distributed in the sodium borosilicate glass. The size of the QDs decreased as the Ag NPs content increased, which in turn depended on the initial amount of Ag2O precursor. The average size of CsPbBr3 QDs was around 4.7, 4.1, 3.4, and 3.3 nm when the initial amount of Ag2O precursor was 0, 0.1, 0.2, and 0.4 mol.%, respectively. The mean size of Ag NPs was 10 nm when the initial amount of Ag2O precursor was 0.1 mol.% (Fig. 7). Notably, the PL of CsPbBr3 QDs/Ag NPs@sodium borosilicate glass (0.1 mol.% of Ag2O precursor) was 2.37 times higher than that of CsPbBr3 QDs@sodium borosilicate glass (0 mol.% of Ag2O precursor). This PL enhancement was facilitated by LSPR in Ag NPs. PL was found to decrease as the amount of Ag NPs increased due to self-adsorption of LSPR. Moreover, the GNC exhibited stability in air, water, and high temperature (Fig. 8) [202]. Javier FONSECA. NPs embedded into glass matrices: GNCs 17 Irradiation Glass matrix −−−−−−−→ h+ + e− (18) Mn+ + ne− → M0 (19) Heat treatment zM0 −−−−−−−−−−→ Mz Fig. 7 (a) TEM image of CsPbBr3 QDs@sodium borosilicate glass. (b) TEM image of CsPbBr3 QDs/Ag NPs@sodium borosilicate glass (0.1 mol.% of Ag2O precursor). Notation: red spheres denote the CsPbBr3 QDs. Reproduced with permission from Ref. [202] (Copyright 2019 John Wiley & Sons). Fig. 8 CsPbBr3 QDs/Ag NPs@sodium borosilicate glass (0.1 mol.% of Ag2O precursor) in water under the irradiation of a 400 nm lamp. Reproduced with permission from Ref. [202] (Copyright 2019 John Wiley & Sons). Metals other than Ag have also been precipitated within glasses. Manzani et al. have precipitated Cu NPs in a glass (Pb2P2O7‒WO3) by conventional melt-quenching method and heat treatment [203]. A mixture of raw materials was melted at 980 °C for 1 h. The molten glass was quenched and annealed at 80 °C for 2 h. To obtain Cu NPs within the glass matrix, the melt-quenched glass was treated at 410 °C during 5, 20, 60, and 120 min. The authors suggest that the Cu NP nuclei were produced during quenching. By heat treatment, the glass allowed the diffusion and aggregation of nuclei which led to the formation of NPs. Therefore, Cu NPs grew by heat treatment at 410 °C for 5 min or more. Specifically, the size of Cu NPs prepared by heat treatment for 120 min was reported to range from 5 to 15 nm. This GNC exhibited high n2, low β, and ultrafast response time [203]. 3.1.3.2 Irradiation-assisted nanoparticle precipitation The mechanism of precipitation of metal NPs by irradiation (and heat treatment) can be expressed as follows: (20) A Mn+-doped glass matrix is irradiated with a femtosecond laser, a nanosecond laser, or X-ray. Electrons (e−) and holes (h+) are generated from the glass matrix due to multiphoton absorption (Eq. (18)). In the area near the focal point, the metal ions (Mn+) are reduced to metal atoms (M0) by trapping the electrons (Eq. (19)). Driven by heat treatment, the M0 atoms diffuse and aggregate to form metal NPs (Mz, where z is the number of M0 aggregated into metal NPs) within the glass matrix (Eq. (20)). The heat treatment refers to the thermal effect of irradiation and/or any other additional thermal approach. Therefore, the precipitation of metal NPs is due to a mixed effect of light and heat. It is worth mentioning that femtosecond lasers generally do not cause heat diffusion in a composite due to the ultrashort time of light-matter interaction. Logically, heat accumulation occurs when femtosecond laser pulses are applied with a high repetition rate. Other NPs such as QDs and NCs tend to precipitate following heat-assisted NPs precipitation and therefore irradiation-assisted NPs precipitation technique is preferably applied to prepare metal NPs. Qiu’s research group has reported the space-selective precipitation of metal NPs by irradiating glass matrices with a focused infrared femtosecond laser [204]. Ag NPs have been precipitated in an Ag+-doped silicate glass (SiO2‒CaO‒Na2O). The glass was prepared by melting the raw materials at 1550 °C for 1 h and subsequently quenching. Ag+-doped silicate glass was irradiated with 400 mW femtosecond laser pulses. Electrons were driven out from the 2p orbital of the NBOs in the SiO4 by the multiphoton absorption after femtosecond laser irradiation. In the area near the focal point, Ag+ ions were reduced to Ag0 by trapping the electrons. The Ag0 atoms aggregated to form NPs after further annealing at 550 °C for 10 min. Therefore, the diffusion of Ag0 was driven by the heat treatment. The focused area turned gray after laser irradiation and yellow after annealing. The size of the Ag NPs was found to range from 1 to 8 nm [176]. Au NPs have also been precipitated inside a Au3+-doped silicate glass (SiO2‒CaO‒Na2O) using femtosecond laser irradiation [177]. A mixture of raw materials was melted 18 Front. Mater. Sci. 2022, 16(3): 220607 at 1550 °C for 1 h and subsequently quenched, forming the Au3+-doped silicate glass. The composite was then irradiated using a femtosecond laser and finally annealed at 550 °C for 30 min. In the focused area, gray spots appeared after irradiation. These turned red after annealing. The photoreduction of Au3+ to Au0 was induced by multiphoton process. Driven by heat treatment, these Au0 atoms aggregated and precipitated into Au NPs. The size of the Au NPs was found to range from 6 to 8 nm. The laser irradiation conditions allowed controlling the size and spatial distribution of the NPs. It should be mentioned that the NPs could also be spaceselectively “dissolved” by femtosecond laser irradiation, and reprecipitated by annealing [177]. Ag and Au NPs have also been co-precipitated within a silicate glass by irradiating femtosecond laser pulses and subsequent annealing at high temperatures [178]. Firstly, a raw material mixture was melted at 1550 °C for 1 h and quenched to prepare the Ag+/Au3+ co-doped silicate glass. The composite was irradiated with a femtosecond laser, releasing electrons. By trapping these electrons, the Ag+ and Au3+ ions were reduced to Ag0 and Au0 atoms, respectively. The subsequent annealing promoted the migration and aggregation of Ag0 and Au0 atoms, thus forming NPs in the glass matrix. The irradiated area exhibited different colors depending on the annealing temperature. While it was yellow after annealing at 400 °C for 30 min, it became red after annealing at 570 °C for 30 min (Fig. 9). Ag NPs were found to emerge earlier than Au NPs during annealing. It is worth mentioning that the precipitation rate of metal NPs is determined by the diffusion rate of metal atoms. Since the radius of the Au atom is bigger than that of the Ag atoms, the Au atom moves more slowly than the Ag atom in the glass Fig. 9 (a) Photograph of the composite after femtosecond laser irradiation and annealing at 400 °C for 30 min. (b) Photograph of the composite after femtosecond laser irradiation and annealing at 570 °C for 30 min. The average laser power was 10 mW. Reproduced with permission from Ref. [178] (Copyright 2006 Elsevier). network. Therefore, the Ag atoms aggregated into the Ag NPs earlier than the Au atoms [178]. The same research group has also precipitated Cu NPs in a borate glass (K2O‒CaO‒B2O3‒CuO) by irradiation with femtosecond laser pulses (Fig. 10) [179]. A mixture of raw materials was melted at 1000 °C for 30 min and subsequently annealed at 300 °C for 2 h. Cu NPs with a size between 3 and 5 nm were precipitated from the CuO-doped borate glass after irradiation with 830 mW femtosecond laser pulses (Fig. 11). A non-linear interaction between the femtosecond laser and the glass sample generated free electrons and holes. These free electrons were trapped by Cu2+ or Cu+ ions, forming Cu0 atoms. Due to the thermal effect of the femtosecond laser, the Cu0 atoms diffused and aggregated to create Cu NPs. Therefore, the Fig. 10 Schematic illustration of femtosecond laser inducing the precipitation of Cu NPs. Reproduced with permission from Ref. [179] (Copyright 2010 Elsevier). Fig. 11 (a) Photograph of the glass sample with a grating pattern after irradiation with a femtosecond laser. (b) Optical microscope photograph of the enlarged grating pattern. (c)(d) Photographs of the diffraction patterns of glass samples before irradiation (panel (c); no diffraction pattern was observed in glass) and after irradiation (panel (d); a diffraction pattern was observed when the grating pattern was irradiated by a 532 nm laser) with a femtosecond laser. Reproduced with permission from Ref. [179] (Copyright 2010 Elsevier). Javier FONSECA. NPs embedded into glass matrices: GNCs precipitation of Cu NPs was due to a mixed effect of light and heat. The area in the vicinity of the laser focal point within the glass sample turned red due to the precipitation of Cu NPs. The red color of the focal regions became deeper and deeper as the irradiation time increased. It was found that femtosecond laser power less than 760 mW did not produce precipitates [179]. Cu NPs have been precipitated within a borosilicate glass by a combination of femtosecond laser irradiation and heat treatment [180]. A Cu2+-doped borosilicate glass (SiO2‒B2O3‒MgO‒Na2O‒Al2O3) was prepared by melting raw materials at 1400 °C for 1 h. The molten composite was quenched and annealed at 400 °C for 12 h. The Cu2+-doped borosilicate glass was irradiated with a femtosecond laser. In doing so, Cu2+ ions were photoreduced to Cu0. The free electrons, which were generated by the non-linear interaction, favored this photoreduction. Subsequently, the composite was treated at 600 °C for 1 h, promoting the aggregation of Cu0 and, consequently, the formation of Cu NPs. These Cu NPs were only observed in the irradiated areas after heat treatment. The mean size of the Cu NPs was 9 nm. A preferential growth of Cu NPs was observed at the bottom of the irradiated region, which was attributed to a selffocusing effect as the beam propagated through the composite [180]. Chen et al. have precipitated Ag NPs in an Ag+-doped soda-lime silicate glass (SiO2‒Na2O‒Al2O3‒CaO) by a combination of X-ray radiation and heat treatment [181]. The glass matrix was prepared by melting a mixture of raw materials at 1450 °C for 4 h. The composite was irradiated with X-ray radiation for 34 min, generating electrons (Eq. (21)). These electrons were trapped by Ag+ ions, producing Ag0 (Eq. (22)). The subsequent heat treatment at 400 °C for 15 min favored the aggregation of Ag0 towards Ag NPs within the glass matrix (Eq. (23)). The size of the Ag NPs ranged between 1 and 5.6 nm with a mean size of (2.8 ± 0.93) nm. X-ray Glass −−−−→ h+ + e− (21) Ag+ + e− → Ag0 (22) zAg0 → Agz (23) The Ag NPs were dissolved by further annealing at 500 °C for 15 min, generating electrons (Eq. (24)), and these electrons recombined with the holes in the glass 19 matrix (Eq. (25)): Agz → zAg+ + ze− (24) h+ + e− → Recombination (25) Therefore, the generation and dissolution of Ag NPs can be controlled by a combination of X-ray radiation and heat treatment [181]. Similarly, Shenget et al. have precipitated Au NPs within an Au3+-doped soda-lime silicate glass (SiO2‒Na2O‒Al2O3‒CaO) by a combination of X-ray radiation and heat treatment [182]. Au NPs have been precipitated in an Au3+-doped silicate glass (SiO2‒Na2O‒CaO) by irradiation (with femtosecond laser, nanosecond laser or X-ray) and heat treatment [205]. A mixture of raw materials was melted at 1550 °C for 4 h and subsequently quenched to prepare the Au3+-doped silicate glass. The composite was irradiated with a femtosecond laser, a nanosecond laser for 4 min, or with X-ray for 4 h. The irradiated composites were then heat treated at 550 °C. The inner electrons of atoms in glass became excited when the composite was irradiated with X-ray. Thus, Au+ ions were reduced to Au0 atoms. Subsequently, the heat treatment promoted the diffusion of Au0 and, therefore, the formation of Au NPs. When the composite was irradiated with a nanosecond laser, the active electrons and holes of NBOs of the glass were excited. Au+ ions captured these free electrons forming Au0 and later NPs in the heat treatment process. Similarly, in the case of femtosecond laser irradiation, the laser beam ionized the glass matrix. Then, Au NPs were created during the heat treatment process. Due to the extremely short time of femtosecond laser irradiation, this approach allowed highly localized laser photons in both time and spatial domains eliminating thermal effects, and non-linear processes. Therefore, femtosecond laser can be more effective than nanosecond laser or X-ray techniques in microprocessing inside glasses. Irradiation was found to reduce the heat treatment temperature required for the precipitation of Au NPs. Whereas Au NPs appeared after heat treatment at 550 °C in irradiated composites, Au NPs only formed at 580 °C in nonirradiated samples. It was found that the mean diameter of the precipitated Au NPs depended on the irradiation step (size of Au NPfemtosecond laser < size of Au NPX-ray < size of Au NPnanosecond laser) [205]. Ion irradiation with very high fluences of protons or He+ ions at energies between 0.1 and 1.5 MeV can also be 20 Front. Mater. Sci. 2022, 16(3): 220607 used to irradiate glass composites and precipitate metal NPs [206‒207]. Herein, NP precipitation occurs during irradiation due to interactions between the metal and irradiation-induced defects. Alternatively, the irradiation of heavy ions (Br−, Si4+ and O2− ions at energies between 6 and 12 MeV) allows the nucleation and growth processes to be completely decoupled, which minimizes the size distribution of the prepared NPs. Valentin et al. have precipitated Cu NPs in a Cu+-doped silicate glass (SiO2‒Na2O‒CaO‒SnO2) by heavy ion irradiation and heat treatment [183]. Ion irradiation initiated the nucleation of Cu0. Cu NPs appeared after post-irradiation heat treatment at 400 °C for 10 min. The growth of NPs was reported to follow the Lifshitz‒Slyozov‒Wagner (LSW) theory of diffusion-limited precipitate ripening [208]. The average size of Cu NPs was found to range from 2 to 8 nm [183]. 3.1.4 Nanoparticles added to glass precursors before melting process NPs can be added to glass precursors before the melting process. This approach allows any type of NP to be incorporated into a glass. However, the NPs can aggregate and dissolve during the melting process to form glass. The melting time and temperature during glass formation are important factors that must be controlled to preserve the structure of the incorporated NPs. The process of dissolution and growth of NPs is called Ostwald ripening. Ostwald ripening is based on the migration of adatoms or mobile species, driven by differences in free energy and local concentrations of adatom [10]. These processes can be thwarted by applying surface engineering to NPs, such as silica coating. Ag NPs have been embedded in Ho3+doped phosphate glass (P2O5‒MgO‒ZnSO4‒Ho2O3) by applying this method [75]. The mixed precursors, including Ag NPs (size < 100 nm), were melted at 1100 °C for 1.5 h and subsequently annealed at 300 °C for 3.5 h. The Ag NPs were homogeneously distributed in the host matrix. The size of the Ag NPs ranged from 2.70 to 9.84 nm with an average diameter of approximately 4.93 nm. The absorption spectra of the composite exhibited 12 bands related to the transitions in the Ho3+ ions. The intensity of PL was influenced by the concentration of Ag NPs. Specifically, the intensity of the green and red emission from the Ho3+ was enhanced by the LSPR of Ag NPs. The effect of LSPR of Ag NPs on the Ho3+ emission was more notable in the composite prepared with low amount of Ho3+ ions. This GNC (AgNPs@Ho3+-doped phosphate glass) is a promising candidate for the development of an efficient solid-state laser medium [75]. Similarly, a borate glass (B2O3‒ZnO‒Na2O) has been doped with NiO NPs by the conventional melt-quenching method [184]. A mixture of raw materials, including NiO NPs with average size of 70 nm, was melted at 1200 °C for 2 h. The molten glass was quenched and subsequently annealed at 400 °C 1 h. This GNC can be used in semiconductor applications [184]. Sm2O3 NPs have been incorporated into a bioborosilicate glass (SiO2‒B2O3‒TeO2) [185]. The silica precursor (98.13 wt.% SiO2 and 1.87 wt.% impurities) was obtained from the waste rice husk. A homogeneous mixture of precursors, which included Sm2O3 NPs, was pre-heated at 400 °C for 1 h and then melted at 1050 °C for 3 h. The molten glass was quenched and annealed at 400 °C for 2 h. While the average size of raw Sm2O3 NPs was around 15‒30 nm, the mean size of embedded Sm2O3 NPs was found to be 71.74 nm. Ostwald ripening can explain the increase in the size of NPs [209]. This composite was suggested to have potential in non-linear optics [185]. In a similar work, Azlina et al. have homogeneously embedded Nd2O3 NPs in a tellurite glass (TeO2‒B2O3‒ZnO). Nd2O3 NPs, TeO2, B2O3, and ZnO were mixed and subsequently melted at 900 °C. The mean size of the raw Nd2O3 NPs was reported to range from 15 to 30 nm. The molted glass was quenched at room temperature producing the GNC (Nd2O3 NPs@tellurite glass) [210]. Likewise, Noorazlan et al. have embedded Er3O2 NPs in the tellurite glass (TeO2‒B2O3‒ZnO) [211]. The GNC was prepared from a mixture of Er2O3 NPs, TeO2, ZnO, and B2O3. The mixture was heated at 400 °C for 30 min, melted at 900 °C for 2 h and finally quenched and annealed at 400 °C for 2 h. The incorporation of Er2O3 NPs within the glass matrix was reported to favor the formation of NBO [211]. Unfortunately, neither the Azlina nor the Noorazlan studies explored the size of the embedded NPs. Halimah et al. have also incorporated Nd2O3 NPs into a tellurite glass (TeO2‒ZnO) [212]. A mixture of precursors, which included Nd2O3 NPs, was heated at 400 °C for 1 h, melted at 830 °C for 1.5 h, and quenched and annealed at 400 °C for 2 h. The average size of embedded Nd2O3 NPs was reported to be approximately 74 nm. Thus, the NPs were suggested to aggregate within the glass matrix [212]. The composite was proposed to be used for linear and NLO applications. Javier FONSECA. NPs embedded into glass matrices: GNCs Nalin’s group has incorporated CdFe2O4 NPs protected with a silica layer into a phosphate glass [213]. First, CdFe2O4 NPs were synthesized by a coprecipitation method. The NPs were coated with a silica layer to protect them during a subsequent melt-quenching process. CdFe2O4 NPs were added to a coacervate, which was previously prepared by mixing solutions of Na(PO3)n and CaCl2. The new mixture was heated to 600 °C for 15 min and then to 1000 °C for 15 min. The CdFe2O4@phosphate glass composite was prepared by quenching the molten glass. Finally, the composite was annealed at 340 °C for 2 h. The embedded CdFe2O4 NPs exhibited a size distribution ranging from 2.1 to 4.2 nm and an average size of 2.8 nm. Considering that the mean size of CdFe2O4 NPs before incorporation was 3.9 nm, the size of the NPs underwent a 29% reduction during melting. The 31P MAS NMR and Raman results showed that the polyphosphate chains were broken and new P−O−Si bonds were formed in the composite [213]. CdFe2O4 NPs coated with a silica layer have also been incorporated in a 60SbPO4‒30ZnO‒ 10PbO glass [186]. The precursors were melted at 1100 °C for 30 min and subsequently quenched, producing the glass 60SbPO4‒30PbO‒10ZnO. The glass was mixed with CdFe2O4 NPs, which were prepared by the coprecipitation method and protected with a silica layer. The mixture was melted at 1100 °C for 7 min and then annealed at 340 °C for 2 h. The CdFe2O4 NPs exhibited an average diameter of 3.5 nm. Considering that the mean size of CdFe2O4 NPs before incorporation was (3.9 ± 0.1) nm, the size of the NPs underwent a 10% reduction during melting. Therefore, the silica layer was 21 able to protect NPs during melting at high temperatures. The matrix acquired thermal stability by the inclusion of CdFe2O4 NPs coated with a silica layer. However, the dispersion of NPs in the glass was not homogeneous [186]. Applying this synthesis approach, the same research group has embedded Fe3−δO4 NPs coated with silica into a phosphate glass (Fig. 12) [214]. Fe3−δO4 NPs were synthesized by thermal decomposition. The NPs were coated by a layer of silica to protect them during the melting. The coacervate gel, which was prepared by mixing Na(PO3)n and CaCl2 solutions, and Fe3−δO4 NPs were melted together at 1000 °C during different times (15, 45, and 75 min). After melting for 15 min, the size distribution of Fe3−δO4 NPs was (19.7 ± 0.6) nm, which was close to that of the as-prepared Fe3−δO4 NPs (20 nm). Silica successfully protected Fe3−δO4 NPs during melting for 15 min. However, the average size of embedded Fe3−δO4 NPs was (9.4 ± 0.6) and (4.3 ± 0.5) nm after melting for 45 and 75 min, respectively. Therefore, the melting time is an important factor to preserve the structure of the NPs embedded in the glass matrix. The incorporation of Fe3−δO4 NPs coated with silica in the phosphate glass favored the formation of P−O−Si bonds, which provided high thermal stability to the GNC. These composites can be used in magneto-optical or ultrasensitive magnetic sensors [214]. Sahar’s research group has prepared GNCs by incorporating nano ferrite oxide (Fe3O4) in glass matrices. This group has incorporated Fe3O4 with a size ranging from 18 to 70 nm in tellurite glass (TeO2‒ZnO) by the melt-quenching technique [187]. A mixture of raw Fig. 12 Schematic illustration of (a) the synthesis of Fe3−δO4@SiO2 NPs and (b) the preparation of the phosphate glasses containing monodisperse Fe3−δO4@SiO2 NPs. Reproduced with permission from Ref. [214] (Copyright 2020 Elsevier). 22 Front. Mater. Sci. 2022, 16(3): 220607 materials, including Fe3O4 NPs, was melted at 850 °C for 20 min and subsequently quenched. Finally, the composite was annealed at 300 °C for 3 h. The GNC containing Fe3O4 NPs exhibited higher thermal stability than the glass without Fe3O4. The GNC was found to be a paramagnetic material. The composite with 2 mol.% Fe3O4 showed a magnetization (M‒H) curve with high saturation magnetic field and a very small hysteresis (Mr = 7.0 × 10−3 emu·g−1, Hc = 1.7 × 10−2 T) [187]. The Fe3O4 NPs have also been incorporated into an Er3+doped tellurite glass (TeO2‒ZnO‒Li2O‒Er2O3) by meltquenching [188]. Again, the mixture of raw materials, which included Fe3O4 NPs with a size ranging from 18 to 70 nm, was melted at 850 °C for 20 min. After quenching, the composite was annealed at 300 °C for 3 h. The Fe3O4 NPs increased the thermal stability of glass. Furthermore, according to the upconversion luminescence spectra, Fe3O4 NPs were found to quench the emission bands of the Er3+ ion [188]. This group has also embedded Fe3O4 NPs into an Er3+-doped phosphate glass (P2O5‒ZnO‒Er2O3) by melt-quenching method [189]. The mixture of raw materials, including Fe3O4 NPs, was melted at 950 °C for 1 h. The molten composite was quenched and annealed at 300 °C for 3 h. Fe3O4 NPs were homogenously distributed within the glass matrix. In the GNC with 1.5 mol.% Fe3O4 NPs, the mean size of the NPs was approximately 31 nm. The average size of NPs in the GNC with 2 mol.% Fe3O4 NPs was approximately 26 nm. The optical bandgap energy for direct and indirect transitions was found to decrease as the content of Fe3O4 NPs increased. It was ascribed to the creation of NBO ions in the glass network. The luminescence was quenched by increasing Fe3O4 NPs concentration. This was attributed to the transfer of energy from the Er3+ ion to the NPs. The GNCs with 1.5‒ 2 mol.% Fe3O4 NPs showed a ferrimagnetic behavior (Fig. 13) [189]. These Fe3O4 NPs-doped composites are promising candidates for magneto-optic devices. 3.2 3.2.1 Sol-gel technique Introduction The sol-gel method is an important wet-chemical approach to synthesize inorganic materials and organic‒ inorganic hybrids from liquid sources. Sol-gel processing of glasses began in the mid-19th century with the studies of Ebelman and Graham [215‒216]. Since then, it has been widely used to prepare glasses [217‒219]. In sol-gel Fig. 13 M‒H curves at room temperature for (a) Fe3O4 NPs and (b) GNCs (Fe3O4 NPs@phosphate glasses) with different concentrations of embedded Fe3O4 NPs. Reproduced with permission from Ref. [189] (Copyright 2015 Elsevier). method, firstly, a suspension of colloidal particles, called sol, is prepared by mixing them in a solvent, generally water, at a pH that prevents precipitation (Fig. 14). Sol particles can also be formed by hydrolyzing a liquid alkoxide precursor in water [220]. The size of the sol particles strongly depends on the pH. It is worth mentioning that the sol, which is a low-viscosity liquid, can be cast into a mold. A gel may be formed by growing an interconnected three-dimensional rigid network from a sol. The gel network may be also formed by the simultaneous hydrolysis and polycondensation of an organometallic precursor [9]. The kinetics of these simultaneous processes depends on the pH, the composition and the concentration of the reagents, the temperature, and pressure [221]. The size of particles and the degree of cross-linking before gelation strongly influence the physical characteristics of the interconnected gel network. Viscosity increases sharply due to gelation. It Javier FONSECA. NPs embedded into glass matrices: GNCs 23 Eq. (26) for a cylindrical pore with radius (r) [223‒227]: pc = − Fig. 14 Schematic illustration of the sol-gel process sequence to prepare NPs embedded into glass matrices. should be noted that the gel must have sufficient strength to resist cracking during thermal treatment. Therefore, before drying, the gel can be maintained immersed in a liquid for a period of time (from hours to days). This aging, also referred to as syneresis, increases the strength of the gel through polycondensation and reprecipitation of the gel network. The pore liquid is generally removed from the interconnected solid gel network by drying between 100 and 180 °C at or near ambient pressure, obtaining a monolith termed as xerogel. When the pore liquid is primarily alcohol-based, the monolith is often referred to as alcogel. The pore liquid can also be evacuated as a gas phase under hypercritical conditions producing a low-density aerogel. The generic term gel is generally applied to xerogels or alcogels, while aerogels are generally referred to as such. A gel is considered as dry when the physisorbed water is completely removed. A dry gel still contains chemisorbed hydroxyls. Heat treatment between 500 and 800 °C desorbs the hydroxyls, resulting in a stabilized gel. The heat treatment of a gel reduces the number of pores and their connectivity, thereby increasing the density of the monolith. The gel is transformed to a dense glass when all pores are eliminated. This process is called densification [11]. The densification temperature ranges from 1000 to 1500 °C, depending considerably on the dimensions of the pore network, the connectivity of the pores, and surface area [222]. As mentioned previously, the drying process must be controlled to avoid the gel to crack. As drying proceeds, capillary pressure (pc) is induced in the pores due to stress arising from solvent evaporation, and pc is expressed by ( ) 2 γ −γ 2γLV cos θ = − SV SL r r (26) where θ is the contact angle, γLV is the interfacial tension between the liquid and the vapor, γSV is the interfacial tension between the solid and the vapor, and γSL is the interfacial tension between the solid and the liquid. Therefore, methods for overcoming stresses during drying include (1) decreasing γLV, (2) increasing r, and (3) increasing θ (or decreasing γSV − γSL ). (1) Methods to decrease γLV. Additives such as surfactants and drying control chemical additives (DCCAs) can be incorporated to lower the γLV of the solvent phase [228‒229]. It is worth mentioning that supercritical drying allows highly porous dry gels to be obtained without fracture [230]. (2) Methods to increase r. Aging can increase the pore size of the wet gels [231]. In addition, this phenomenon strengthens the gel framework, preventing fracture during drying and bloating during sintering [232‒233]. It is worth mentioning that gels with small pores (typically around 5 nm or so) can be slowly dried without fracture [234‒237]. Interestingly, particles can be incorporated into precursor solution to reduce the pc [227]. (3) Methods to increase θ (or decrease γSV − γSL ). The surface of the wet gel can be modified to form a hydrophobic surface [238]. The sol-gel approach is a melt-free process. Therefore, it requires lower processing temperature than traditional glass melting methods. This method is versatile and flexible, specifically, it has potential to increase the concentration of reinforcements and broaden the compositional range of glasses. In addition, this technique provides materials of high purity and homogeneity. Glasses with high melting temperatures, high crystallization tendencies, and compositions within the stable or metastable liquid‒liquid immiscibility region are preferably prepared by the sol-gel method [239‒240]. Thus, this approach is particularly attractive for the development of GNCs that are difficult to prepare by melt-quenching and vapor phase methods. The NP precursors can be initially mixed with glass precursors. Thus, glass matrix and NPs can be synthesized concomitantly. In doing so, the reinforcements are usually homogeneously distributed within the resulting glass matrix. An alternative to such an approach is a sol-gel method for glass synthesis combined with impregnation with the NP or NP precursor, generally 24 Front. Mater. Sci. 2022, 16(3): 220607 followed by heat treatment. The following sections of this review are intended to provide an overview of the main sol-gel approaches for GNC synthesis. Table 3 shows examples of GNCs prepared by sol-gel techniques [14‒15,241‒257]. 3.2.2 Concomitant synthesis of glass matrix and nanoparticles Schubert’s research group have prepared composites containing uniform metal NPs, homogeneously dispersed in SiO2 matrices by a three-step procedure [241,258]: (1) An alkoxysilane of the type (RO)3Si(CH2)nA (A = coordinating organic group = NH2, NHCH2CH2NH2, CN), a metal salt (AgNO3, AgOAc, Cd(NO3)2, Co(OAc)2, Cu(OAc)2, Ni(OAc)2, Pd(acac)2, or Pt(acac)2), and, optionally, Si(OR)4 were processed by the sol-gel method, producing a gel with composition [O3/2Si(CH2)nA]yMXm· xSiO2. The metal loading was adjusted by adding the Si(OR)4. (2) The polycondensates were calcined in air, and thereby the metal oxide@SiO2 composites (MOz@(x + y)SiO2) (MOz = Ag2O, CdO, CoO, CuO, NiO, PdO or PtO) were formed. The oxidation temperature ensured the complete oxidation of all organic components, without causing excessive sintering of NPs. (3) The metal oxide particles were reduced by H2, generating the nanocomposites M@(x + y)SiO2 (M = Ag, Co, Cu, Ni, Pd, or Pt). Therefore, reduction of the metal precursor with H2 at high temperature was required to create the NPs. The metal particle size did not change significantly during the reduction step [241,258]. The size of the embedded metal NPs was found to depend on the metal, the metal loading, the chemical composition of both the metal precursor and the organo(alkoxy)silane, and the conditions for calcination [242,259‒260]. Moreover, the textural properties (surface area, porosity, etc.) of the prepared composites have been reported to depend on the metal precursor [243,261]. PdNi alloy@silica glass and CuNi alloy@silica glass composites have also been prepared by this approach [262‒263]. In brief, Schubert’s approach tethers metal precursors to the matrix during solgel processing to embed metal NPs into silica glass. This technique makes possible to prepare MOx@SiO2 or M@SiO2 nanocomposites with a controllable size of NPs. In addition, the method has been extended to more complex systems such as complex hybrid organic — “multi inorganic” particles SiO2@PDCL-ZrO2@ (Ag@ SiO2) (PDCL = poly(N-dicarbazolyl-lysine)) [264]. Crystalline nanoclusters of Fe2P, RuP, Co2P, Rh2P, Ni2P, Pd5P2, and PtP2 have been formed in silica xerogel by a similar approach [244]. The general synthesis of metal phosphide@silica xerogel nanocomposites also required three steps: (1) Preparation of single-source transition metal complexes that served as precursors to metal phosphide nanoclusters. The metal complexes contained bifunctional phosphine ligands possessing alkoxysilyl (Si-(OR)3) functional groups. (2) Covalent incorporation of the precursors into the silica xerogel matrixes as they were being formed. The silica xerogel host matrix was formed by sol-gel approach. Therefore, the precursors exhibited adequate solubility and chemical stability to survive sol-gel process without undergoing chemical degradation. (3) Heat treatment under reducing conditions to convert the doped xerogel host matrix to a metal phosphide@ silica xerogel nanocomposite. The nanoclusters of Fe2P, RuP, Co2P, Rh2P, Ni2P, Pd5P2, or PtP2 were highly dispersed throughout the xerogel matrix. The size of the NPs ranged from 2.0 to 11.3 nm [244]. Nucleation and growth of NPs were suggested to be driven by diffusion, similar to the processes observed in the preparation of the GNCs (CdS@silica or CdSe@silica) by melt-quenching from supersaturated glasses at high temperature [164]. The synthesis procedure described above, based on the covalent incorporation of suitable precursors into silica xerogel matrices and reduction at high temperature, has also been followed in the preparation of Co3C@silica xerogel, Ge@silica xerogel, Os@silica xerogel, and PtSn@silica xerogel composites [245‒247]. CdS QDs have been embedded into a sodium borosilicate glass matrix via sol-gel method [248]. A cadmium-doped sodium borosilicate gel was prepared by mixing the precursors (sodium silicate (Na2SiO3), boric acid (H3BO3), and cadmium sulfate (CdSO4)) in an aqueous solution at low temperature (60‒80 °C). Sulfurcontaining complexes were dissolved in the gel to perform Cd2+ sulphidation during gel stabilization at temperatures ranging from 300 to 500 °C. After quenching the gel from 750 °C to room temperature, the crystalline quality of CdS QDs was improved by annealing the glass. The size of CdS QDs was reported to range from 4 to 20 nm depending on the annealing temperature and the initial Silica glass Silica glass Silica glass Silica glass Silica glass Ag NPs@silica glass (Ag·4SiO2) Ag NPs@silica glass (Ag·12SiO2) Co NPs@silica glass (Co·SiO2) Cu NPs@silica glass (Cu·SiO2) )a) Silica glass Silica glass Silica glass Silica glass Ni NPs@silica glass (Ni·3SiO2)a) Ni NPs@silica glass (Ni·5.5SiO2)a) Ni NPs@silica glass (Ni·10SiO2)a) )a) RuP NCs Co2P NCs Silica xerogel Silica xerogel Silica xerogel Silica xerogel RuP NCs@silica xerogel Co2P NCs@silica xerogel Rh2P NCs@silica xerogel Ni2P NCs@silica xerogel Ni2P NCs Rh2P NCs Fe2P NCs Ni NPs Ni NPs Ni NPs Silica glass Ni NPs@silica glassb)g) Silica xerogel Silica glass Ni NPs@silica Ni NPs Fe2P NCs@silica xerogel Silica glass glassb)f) Ni NPs@silica Pt NPs Pt NPs Pt NPs Pt NPs Pt NPs Pd NPs Pd NPs Pd NPs Pd NPs Pd NPs Pd NPs Pd NPs Ni NPs Ni NPs Ni NPs Ni NPs Ni NPs Ni NPs Cu NPs Co NPs Ag NPs Ag NPs Reinforcement Ni NPs@silica glassb)h) Silica glass glassb)e) Pt NPs@silica Silica glass Silica glass Pt NPs@silica glass (Pt·32SiO2) Silica glass Silica glass Pd NPs@silica glass (Pd·15SiO2) Pt NPs@silica glassd) Silica glass Pd NPs@silica glass (Pd·32SiO2) Pt NPs@silica glassc) Silica glass Pd NPs@silica glass (Pd·22SiO2) Silica glass Silica glass Pd NPs@silica glass (Pd·12SiO2) Silica glass Silica glass Pd NPs@silica glass (Pd·7SiO2) Pt NPs@silica glassb) Silica glass Pd NPs@silica glass (Pd·4.5SiO2) glassa) Silica glass Pd NPs@silica glass (Pd·2SiO2) Ni NPs@silica glass (Ni·33SiO2 Silica glass Ni NPs@silica glass (Ni·2SiO2)a) Ni NPs@silica glass (Ni·SiO2 Host matrix GNCs prepared by sol-gel techniques [14,15,241–257] Composite Table 3 (2.6) (2.0) (5.0) (4.7) (4.6) (8.5) (8.2) (8.5) (8.5) (5.0) (4.0) (3.5) (7.1) 0.8‒4.2 (2.5) 2.8‒5.2 (3.8) 1.3‒3.7 (2.4) ‒ ‒ 1.8‒4.2 (2.8) ‒ 1.8‒4.2 (3.0) 2.5‒12.4 (5.9) Bimodal particle distribution: 7.5‒12.4, 27.5‒72.4 Bimodal particle distribution: 2.5‒22.4, 62.5‒102.4 Bimodal particle distribution: 2.5‒12.4, 32.5‒57.4 12.5‒47.4 (22.9) 37.5‒62.4 (50.3) 1.5‒7.4 (3.9) 11.0‒24.9 (17.4) 9.0‒30.9 (19.5) ‒ Size of NPs (min‒max (mean))/nm 6.83 8.79 4.96 8.23 6.64 10 24 9 4 21.5 24.3 24.8 17 7.93 9.36 5.52 7.18 10.84 18.12 26.03 44.81 2.65 6.99 14.01 20.05 31.08 43.7 46.17 47.51 28.7 10.22 Loading of NPs/wt.% Concomitant synthesis of glass matrix and NPs: tethering metal precursors to the matrix during sol-gel processing Concomitant synthesis of glass matrix and NPs: tethering metal precursors to the matrix during sol-gel processing Concomitant synthesis of glass matrix and NPs: tethering metal precursors to the matrix during sol-gel processing Concomitant synthesis of glass matrix and NPs: tethering metal precursors to the matrix during sol-gel processing Sol-gel technique ‒ ‒ ‒ ‒ Application [244] [243] [242] [241] Ref. Javier FONSECA. NPs embedded into glass matrices: GNCs 25 (32.63) 1.5‒5 (2.7) (30.28) (23) Sb NPs Cu NPs Cu2In NPs In2O3 NCs In2O3 NCs In2O3 NCs@silica glass Au NPs@sodium borosilicate glass Sodium borosilicate glass (75 wt.% SiO2, 20 wt.% B2O3, and 5 wt.% Na2O) Cu3.8Ni@sodium borosilicate glass Sodium borosilicate glass (72.88 wt.% SiO2, 21.38 wt.% B2O3, and 5.74 wt.% Na2O) Au NP/Ni NP@sodium Sodium borosilicate glass borosilicate glass (75 wt.% SiO2, 20 wt.% B2O3, and 5 wt.% Na2O) 2.82‒9.97 (5.48) 15‒40 (27.5) 8 nm Au NPs, 2 nm Ni NPs Au NPs Cu3.8Ni alloy NPs Au NPs and Ni NPs (54) (4.3) CdS QDs Silica glass (5.7) 2‒14 (6) CdS QDs PtSn alloy NPs 1‒7 (2.1) (15.4) Silica xerogel PtSn@silica xerogel Os NPs 2.5‒14.5 (6.7) CdS QDs Silica xerogel Os NPs@silica xerogel Ge NCs 10‒46 (25) (18.4) Silica xerogel Ge NCs@silica xerogel Co3C NCs (11.3) (4) Size of NPs (min‒max (mean))/nm CdS QDs Silica xerogel Co3C NCs@silica xerogel Pd5P2 NCs PtP2 NCs Reinforcement Sodium borosilicate glass (33.33 wt.% SiO2, 33.33 wt.% B2O3, and 33.33 wt.% Na2O) CdS QDs@sodium borosilicate Sodium borosilicate glass j) glass (33.33 wt.% SiO2, 33.33 wt.% B2O3, and 33.33 wt.% Na2O) CdS QDs@sodium borosilicate Sodium borosilicate glass k) glass (33.33 wt.% SiO2, 33.33 wt.% B2O3, and 33.33 wt.% Na2O) CdS QDs@sodium borosilicate Sodium borosilicate glass glassl) (33.33 wt.% SiO2, 33.33 wt.% B2O3, and 33.33 wt.% Na2O) Sb NPs@sodium borosilicate glass Sodium borosilicate glass (72.88 wt.% SiO2, 21.38 wt.% B2O3, and 5.74 wt.% Na2O) Cu NPs@sodium borosilicate glass Sodium borosilicate glass (72.88 wt.% SiO2, 21.38 wt.% B2O3, and 5.74 wt.% Na2O) Cu2In NPs@sodium borosilicate Sodium borosilicate glass glass (72.88 wt.% SiO2, 21.38 wt.% B2O3, and 5.74 wt.% Na2O) In2O3 NCs@silica glass Silica glass Silica xerogel Silica xerogel Pd5P2 NCs@silica xerogel PtP2 NCs@silica xerogel CdS QDs@sodium borosilicate glassi) Host matrix Composite 2 wt.% Au, 1 wt.% Ni 1.5 0.25 3 1 1.5 1.25 1.5 ‒ ‒ ‒ ‒ ‒ 0.2‒1 ‒ ‒ 25.84 2.77 Loading of NPs/wt.% ‒ ‒ ‒ Application NLO devices: χ(3)(800 nm) = 3.32 × 10−10 esu NLO devices: χ(3)(800 nm) = 2.41 × 10−11 esu NLO devices: χ(3)(800 nm) = 4.85 × 10−11 esu Concomitant synthesis of glass matrix and NPs: sol-gel method combined with heat treatment in a suitable atmosphere Concomitant synthesis of glass matrix and NPs: sol-gel method combined with heat treatment in a suitable atmosphere Concomitant synthesis of glass matrix and NPs: sol-gel method combined with heat treatment in a suitable atmosphere NLO devices: χ(3)(800 nm) = 2.31 × 10−12 esu NLO devices: χ(3)(800 nm) = 1.7 × 10−14 esu NLO devices: χ(3)(800 nm) = 4.92 × 10−11 esu Concomitant synthesis of glass matrix Photocatalytic and NPs: sol-gel method combined with degradation of heat treatment in a suitable atmosphere organic pollutants Concomitant synthesis of glass matrix and NPs: sol-gel method combined with heat treatment in a suitable atmosphere Concomitant synthesis of glass matrix and NPs: sol-gel method combined with heat treatment in a suitable atmosphere Concomitant synthesis of glass matrix and NPs: sol-gel method combined with heat treatment in a suitable atmosphere Concomitant synthesis of glass matrix Solar concentration; and NPs lasers Concomitant synthesis of glass matrix and NPs: tethering metal precursors to the matrix during sol-gel processing Concomitant synthesis of glass matrix and NPs: tethering metal precursors to the matrix during sol-gel processing Concomitant synthesis of glass matrix and NPs: tethering metal precursors to the matrix during sol-gel processing Sol-gel technique [254] [253] [252] [251] [15] [250] [249] [248] [247] [246] [245] Ref. (continued) 26 Front. Mater. Sci. 2022, 16(3): 220607 Sm2O3 NPs Sm2O3 NPs Au NPs AuAg alloy NPs Au NPs and Ag NPs CuO NCs Cu2O NCs (4.1) (2.7) 10‒30 5.6‒51.9 (16) 3.7‒63 (22) 1‒5 (2.24) 1‒6 (3.12) 9‒34 (18.8) 22.3 nm Au NPs, 22.3 nm Cu NPs Au NPs and Cu NPs Cu NPs 18.08 nm Au NPs, 18.08 nm Cu NPs Au NPs and Cu NPs 12.89 nm Au NPs, 12.89 nm Cu NPs 7 nm Au NPs, 7 nm NiO NPs Au NPs and NiO NPs Au NPs and Cu NPs Size of NPs (min‒max (mean))/nm Reinforcement 4.44 5.12 ‒ ‒ ‒ 2 2 2 0.25 wt.% Au, 0.75 wt.% Cu 0.25 wt.% Au, 0.75 wt.% Cu 0.25 wt.% Au, 0.75 wt.% Cu 2 wt.% Au, 1 wt.% NiO Loading of NPs/wt.% Art NLO devices: χ(3)(800 nm) = 1.81 × 10−13 esu NLO devices: χ(3)(800 nm) = 3.1 × 10−12 esu NLO devices: χ(3)(800 nm) = 5.4 × 10−12 esu NLO devices: χ(3)(800 nm) = 4.4 × 10−12 esu NLO devices: χ(3)(800 nm) = 6.4 × 10−14 esu NLO devices: χ(3)(800 nm) = 1.6 × 10−14 esu NLO devices: χ(3)(800 nm) = 2.6 × 10−14 esu Art Application NLO devices: β(532 nm) = −1.37 × 10−10 m·W−1 Sol-gel technique combined with NP or High-density optical NP precursors impregnation: sol-gel storage, undersea method and loading of NP precursors communication, and color displays Sol-gel technique combined with NP or NP precursors impregnation: sol-gel method and loading of NP precursors Concomitant synthesis of glass matrix and NPs: sol-gel method combined with heat treatment in a suitable atmosphere Concomitant synthesis of glass matrix and NPs: sol-gel method combined with heat treatment in a suitable atmosphere Concomitant synthesis of glass matrix and NPs: sol-gel method combined with heat treatment in a suitable atmosphere Concomitant synthesis of glass matrix and NPs: sol-gel method combined with heat treatment in a suitable atmosphere Concomitant synthesis of glass matrix and NPs: sol-gel method combined with heat treatment in a suitable atmosphere Concomitant synthesis of glass matrix and NPs: sol-gel method combined with heat treatment in a suitable atmosphere Concomitant synthesis of glass matrix and NPs: sol-gel method combined with heat treatment in a suitable atmosphere Sol-gel technique combined with NP or NP precursors impregnation: sol-gel method and loading of NPs Sol-gel technique combined with NP or NP precursors impregnation: sol-gel method and loading of NPs Sol-gel technique combined with NP or NP precursors impregnation: sol-gel method and loading of NPs Sol-gel technique [257] [256] [156] [14] [255] Ref. a) A = 3-(2-aminoethylamino)propyltriethoxysilane (AEAPTS); b) A = aminopropyltriethoxysilane (APS); c) A = (3-trimethoxysilylpropyl) diethylenetriamine (TAS); d) A = 2-(trimethoxysilyl-ethyl)pyridine (PyS); e) nickel precursor = Ni(NO3)2; f) nickel precursor = Ni(OAc)2; g) nickel precursor = NiCl2; h) nickel precursor = Ni(acac)2; i) 0.3 mol.% Cd precursor; j) 0.03 mol.% Cd precursor; k) 0.017 mol.% Cd precursor, 4.5 h at 540 °C; l) 0.010 mol.% Cd precursor, 4.5 h at 540 °C; m) red; n) green; o) blue. Sm2O3 NPs@sodium borosilicate glass Borosilicate glass (85.17 wt.% SiO2, 14.83 wt.% B2O3) Sodium borosilicate glass (77.80 wt.% SiO2, 13.54 wt.% B2O3, and 8.65 wt.% Na2O) Silica glass Au NPs@silica glass Sm2O3 NPs@borosilicate glass Silica glass Sodium borosilicate glass (75 wt.% SiO2, 20 wt.% B2O3, and 5 wt.% Na2O) Sodium borosilicate glass (96.7 wt.% SiO2, 2.7 wt.% B2O3, and 0.6 wt.% Na2O) Sodium borosilicate glass (96.7 wt.% SiO2, 2.7 wt.% B2O3, and 0.6 wt.% Na2O) Sodium borosilicate glass (96.7 wt.% SiO2, 2.7 wt.% B2O3, and 0.6 wt.% Na2O) Sodium borosilicate glass (75 wt.% SiO2, 20 wt.% B2O3, and 5 wt.% Na2O) Sodium borosilicate glass (75 wt.% SiO2, 20 wt.% B2O3, and 5 wt.% Na2O) Sodium borosilicate glass (75 wt.% SiO2, 20 wt.% B2O3, and 5 wt.% Na2O) Silica glass Host matrix AuAg@silica glass Au NPs/Ag NPs@silica glass CuO NCs@borosilicate glasso) Cu2O NCs@borosilicate glassn) Cu NPs@borosilicate glassm) Au NPs/Cu NPs@sodium borosilicate glass (3) Au NPs/Cu NPs@sodium borosilicate glass (2) Au NPs/Cu NPs@sodium borosilicate glass (1) Au NP/NiO NP@sodium borosilicate glass Composite (continued) Javier FONSECA. NPs embedded into glass matrices: GNCs 27 28 Front. Mater. Sci. 2022, 16(3): 220607 concentration of cadmium precursor. Alternatively, H2S gas flow was also used to achieve Cd2+ sulphidation. In this case, the size of CdS QDs depended on the temperature and time of exposure to H2S and the concentration of Cd2+ [248]. The most immediate applications of this composite are as solar concentrators and as an active medium in tunable lasers [265]. Xiang’s group has combined the sol-gel method with heat treatment in suitable atmospheres, which allows uniform distribution of reinforcements and controls the reinforcement concentrations. In this approach, the dry gel is generally treated in three atmospheres: (1) an oxidizing atmosphere, eliminating the organic residues of the precursors and producing an aerogel; (2) a reducing atmosphere, forming the NPs; and (3) an inert atmosphere, densifying the stabilized gel and producing the final composite (NPs@glass). It is worth mentioning that this method requires low temperature and, therefore, is a low-cost synthesis route. Many GNCs have been prepared following this approach [266]: (a) Sb NPs have been embedded in a sodium borosilicate glass matrix (Fig. 15). Antimony trichloride (SbCl3), tetraethoxysilane (TEOS), H3BO3, and metallic sodium were used as Sb ion source, SiO2 precursor, boron source, and sodium source, respectively. Gelation and Fig. 15 (a) TEM image of Sb NPs embedded in sodium borosilicate glass. (b) Size distribution of Sb NPs in the glass matrix. (c) HRTEM of Sb NPs@sodium borosilicate glass. (d) SAED image of Sb NPs@sodium borosilicate glass. Reproduced with permission from Ref. [249] (Copyright 2014 Elsevier). aging were performed at 80 °C for two weeks. Subsequently, the dry gel was heated in three atmospheres: (1) O2 at 450 °C for 10 h, producing an aerogel; (2) dry H2 at 450 °C for 10 h, obtaining the Sb NPs; and (3) N2 at 600 °C for 10 h, creating the final composite (Sb NPs@sodium borosilicate glass). Sb NPs were reported to have a spherical shape and an average size of approximately 32.63 nm. At the wavelength of 800 nm, β, n2, and χ(3) of this GNC were −1.71 × 10−10 m·W−1, 8.15 × 10−16 m2·W−1, and 4.85 × 10−11 esu, respectively. Therefore, this Sb NPs@sodium borosilicate glass composite has potential for NLO devices [249]. (b) Sodium borosilicate glass has been doped with Cu NPs. Copper nitrate (Cu(NO3)2), TEOS, H3BO3, and metallic sodium were used as Cu ion source, SiO2 precursor, boron source, and sodium source, respectively. The sol was gelled and aged at 80 °C for 2 weeks. To remove the organic substances and decompose Cu(NO3)2, the wet gel was dried in O2 at 450 °C for 10 h. Then, the dry gel was exposed to dry H2 at 450 °C for 10 h in order to convert CuO to Cu. Finally, the stabilized gel was densified in N2 at 600 °C for 10 h. The size of the Cu NPs ranged from 1.5 to 5 nm with a mean particle size of 2.7 nm. At the wavelength of 800 nm, β, n2, and χ(3) were found to be 2.10 × 10−11 m·W−1, 6.42 × 10−17 m2·W−1, and 2.41 × 10−11 esu, respectively [250]. (c) Cu2In NPs has been incorporated into a sodium borosilicate glass. Cu(NO3)2, indium nitrate (In(NO3)3), TEOS, H3BO3, and metallic sodium were used as Cu ion source, In ion source, SiO2 precursor, boron source, and sodium source, respectively. The precursors were mixed, gelled for 2‒3 d, and dried at 80 °C for 2‒3 weeks. The dry gel was heated in O2 atmosphere at 450 °C for 10 h. Then, the aerogel was treated with dry H2 at 450 °C for 10 h. Finally, the aerogel was heated in N2 atmosphere at 600 °C for 10 h. The size distribution of Cu2In NPs was reported to range from 15 to 50 nm with an average size of approximately 30.28 nm. Cu2In NPs exhibited hexagonal structure and spherical shape. At the wavelength of 800 nm, β, n2, and χ(3) of these NPs embedded in sodium borosilicate glass were 6.96 × 10−10 m·W−1, −5.99 × 10−16 m2·W−1, and 3.32 × 10−10 esu, respectively [15]. (d) In2O3 semiconductor NCs have been embedded into silica glass. In(NO3)3, and TEOS were used as In ion source, and SiO2 precursor, respectively. The precursor solution gelled at room temperature for 7 d. The gel was aged at 150 °C for 3 to 4 weeks. To eliminate organic Javier FONSECA. NPs embedded into glass matrices: GNCs residues, the gel was heated in O2 atmosphere at 600 °C for 5 h, obtaining the In2O3@SiO2 composite. The mean size of the obtained In2O3 NCs was found to increase from 23 to 54 nm when the doping of the glass matrix increased from 1 to 3 wt.% of In2O3. The photodegradations of rhodamine (RhB) under UV irradiation in the presence of In2O3 (1 wt.%)@SiO2 and In2O3 (3 wt.%)@ SiO2 composites for 120 min were ~35.5% and ~61.6%, respectively. The photodegradation efficiency of silica glass was significantly improved by embedding In2O3 NCs. Therefore, In2O3@SiO2 composites are promising photocatalysts for the degradation of organic pollutants [251]. (e) Sodium borosilicate glass has been doped with Au NPs (Fig. 16). Hydrogen tetrachloroaurate (HAuCl4), TEOS, H3BO3, and metallic sodium were used as Au ion source, SiO2 precursor, boron source, and sodium source, respectively. Within a week, a wet gel was formed from the mixed solution. The wet gel was dried at 80 °C for about 3 weeks to obtain the dry gel. To remove the residual organic solvent and to decompose HAuCl4, the dry gel was heated in O2 atmosphere at 450 °C for 5 h. Then, it was exposed to H2 at 450 °C for 5 h, forming metallic Au. Finally, the stabilized gel was densified in N2 atmosphere at 600 °C for 2 h, producing the Fig. 16 (a) TEM image of Au NPs embedded in sodium borosilicate glass. (b) Size distribution of Au NPs in the glass matrix. (c) HRTEM of Au NPs@sodium borosilicate glass. (d) SAED image of Au NPs@sodium borosilicate glass. Reproduced with permission from Ref. [252] (Copyright 2015 Springer Nature). 29 AuNPs@sodium borosilicate glass composite. The size of embedded Au NPs was found to range from 2.82 to 9.97 nm, with an average size of 5.48 nm. At the wavelength of 800 nm, β, n2, and χ(3) were −6.5 × 10−14 m·W−1, 3.0 × 10−20 m2·W−1, and 1.7 × 10−14 esu, respectively [252]. (f) Cu3.8Ni NCs have been embedded in a sodium borosilicate glass matrix. Nickel nitrate (Ni(NO3)2), Cu(NO3)2, TEOS, H3BO3, and metallic sodium were used as Ni ion source, Cu ion source, SiO2 precursor, boron source, and sodium source, respectively. After preparation and gelation of the sol, the gel was heated in H2 atmosphere at 600 °C, producing the composite. XPS analysis indicated that the obtained nanocrystal was Cu3.8Ni alloy. The size of Cu3.8Ni NCs was found to range from 15 to 40 nm with an average size of 27.5 nm. At the wavelength of 800 nm, β, n2, and χ(3) were (3.73 ± 0.02) × 10−10 m·W−1, (1.07 ± 0.08) × 10−15 m2·W−1, and (4.92 ± 0.05) × 10−11 esu, respectively. χ(3) was larger than those of the corresponding monometallic NPs [253]. (g) Au and NiO NPs or Au and Ni NPs have been incorporated in sodium borosilicate glass. HAuCl4 and Ni(CH3COO)2 were used as Au ion source and Ni ion source, respectively. The Au NPs/NiO NPs@sodium borosilicate glass composite was synthesized by treating the gel in O2 atmosphere at 450 °C for 10 h and N2 atmosphere at 600 °C for 2 h. The Au NPs/Ni NPs@sodium borosilicate glass composite was fabricated when the gel was exposed to O2 at 450 °C for 10 h, H2 atmosphere at 450 °C for 10 h, and finally, N2 atmosphere at 600 °C for 2 h. The NPs of the Au NPs/NiO NPs@sodium borosilicate glass composite exhibited a size of 2 and 8 nm. The smallest and largest NPs were suggested to be NiO and Au, respectively. In Au NPs/Ni NPs@sodium borosilicate glass composite, the size of both NPs was found to be around 7 nm. They exhibited superior optical non-linearities compared to Au NPs@sodium borosilicate glass composite. At the wavelength of 800 nm, χ(3) of Au NPs/NiO NPs@sodium borosilicate glass composite and Au NPs/Ni NPs@sodium borosilicate composite was reported to be 1.81 × 10−13 and 2.31 × 10−12 esu, respectively [254]. (h) Au and Cu NPs have been embedded in sodium borosilicate glass. Cu(NO3)2, HAuCl4, TEOS, H3BO3, and metallic sodium were used as Cu ion source, Au ion source, SiO2 precursor, boron source, and sodium source, respectively. Gelation, aging and drying of the mixture 30 Front. Mater. Sci. 2022, 16(3): 220607 was carried out by heating at 100 °C for 3 weeks. Then, the dry gel was exposed to a H2 atmosphere at 450 °C for 10, 20, and 30 h, producing stabilized gels referred to as (1), (2), and (3), respectively. Finally, the stabilized gels were densified in a N2 atmosphere at 600 °C. No AuNi alloy was observed. It was suggested that the Au NPs migrated toward the core of the matrix and the Cu NPs coated the surface of the matrix. The mean size of the NPs embedded in the glass matrix was found to be 12.89, 18.08, and 22.30 nm for GNC (1), GNC (2), and GNC (3), respectively. At the wavelength of 800 nm, χ(3) values of GNC (1), GNC (2), and GNC (3) were 3.1 × 10−12, 5.4 × 10−12, and 4.4 × 10−12 esu, respectively. Thus, the NPs improved the optical non-linearities of the glass matrix [255]. (i) Red, green and blue glass composites have been prepared by embedding Cu NPs, Cu2O NCs and CuO NCs into sodium borosilicate glass, respectively (Fig. 17). Cu(NO3)3, TEOS, H3BO3, and metallic sodium were used as Cu ion source, SiO2 precursor, boron source, and sodium source, respectively. The mixture solution gelled at 80 °C for 3 weeks. The wet gel was exposed to an O2 atmosphere at 400 °C for 10 h. In doing so, the organic matter was removed, and the copper nitrate decomposed. The dry gel was heated in H2 atmosphere at 600 °C for 10, 5, and 0 h, forming Cu NPs, Cu2O NCs, and CuO NCs, respectively. Therefore, depending on exposure time in a reducing atmosphere, the valence state of copper was 0 (Cu0), +1 (Cu1+) or +2 (Cu2+). Finally, the stabilized gel was densified in N2 atmosphere at 600 °C for 5 h, producing the composites. The red, green, and blue colors of the composites were found to correspond to Cu (Cu0), Cu2O (Cu1+), and CuO (Cu2+) NCs embedded in the glass matrix, respectively. The size of these three different crystals in the glass matrices ranged from 9 to 34 nm, 1 to 6 nm, and 1 to 5 nm, respectively. The average size of the crystals was found to be 18.8, 3.12, and 2.24 nm, respectively. At the wavelength of 800 nm, χ(3) of the red, green, and blue composites was 6.4 × 10−14, 1.6 × 10−14, and 2.6×10−14 esu, respectively [14]. Yanes et al. have prepared an YF3 NC@Ln3+-doped silica glass composite, where Ln3+ = Eu3+ or Sm3+, and an YF3 NC@Yb3+, Tm3+ co-doped silica glass composite by sol-gel method [267]. An initial homogeneous solution containing yttrium acetate (Y(CH3COO)3), TEOS and europium acetate (Eu(CH3COO)3) or samarium nitrate (Sm(NO3)3) gelled at 35 °C for several days. The gel was treated in air at 675 °C to precipitate YF3 NCs, thus forming the YF3 NC@Ln3+-doped silica glass composite. The same procedure was used to prepare YF3 NC@Yb3+, Tm3+ co-doped silica glass composite. In this case, the source of Yb3+ was ytterbium acetate (Yb(CH3COO)3). YF3 NCs of 11 nm were found to precipitate in the glass matrices. YF3 is a quantum efficient host lattice for RE ions due to its wide band gap and the possibility of replacing Y3+ sites by trivalent RE ions with similar ionic radii and valence state [268‒269]. YF3 NC@Ln3+-doped silica glass composite spectra revealed a significant partition of Ln3+ into the precipitated YF3 NCs. Moreover, YF3 NC@Yb3+, Tm3+ co-doped silica glass composite achieved bright and efficient infrared to UV and visible up-conversion. It is worth mentioning that the UV up-conversion emissions improved with increasing Yb3+ concentration. YF3 NC@Yb3+, Tm3+ co-doped silica glass composite opens the door to develop shortwavelength solid-state lasers for applications in integrated photonic devices [267]. Recently, Kalwarczyk et al. have reported the synthesis of Au NPs@silica glass composite by sol-gel method. HAuCl4, TEOS, and L-ascorbic acid were used as Au ion source, SiO2 precursor, and reducing agent, respectively. L-ascorbic acid was applied in an amount sufficient for partial reduction of Au3+ ions to Au+. The Au+ ions were suggested to be encapsulated within micelles, preventing their reduction to Au0 and facilitating their homogeneous distribution in the gel. The micelles were made of cationic surfactants (benzyldimethylhexadecylammonium chloride (BDAC) or/and cetyltrimethylammonium bromide (CTAB)). The reduction of Au+ to Au0 (Au NPs) was proposed to occur when micelles break during the hydrolysis of TEOS. Therefore, Au NPs were formed through in situ nucleation and growth during the gelation and/or drying steps. It should be noted that the growth of the Au NPs within the silica matrix occurred at only 35 °C for 7‒8 d. Moreover, the Au NP@silica glass composite did not require any post-synthetic treatment such as heat. This GNC was found to contain 0.027 wt.% of Au. The average size of the embedded Au NPs was (6.9 ± 0.3) nm. Au NP@silica glass composite was very stable. Embedded Au NPs did not aggregate or disintegrate for one year. Despite small amount of Au NPs, the composite exhibited strong NLO properties, and was capable of effective third-harmonic generation of ultrashort near-IR laser pulse. Therefore, this GNC can be potentially employed as an IR-UV converter [270]. Javier FONSECA. NPs embedded into glass matrices: GNCs 31 Fig. 17 TEM images of (a) red GNC, (b) green GNC, and (c) blue GNC. HRTEM images of (d) red GNC, (e) green GNC, and (f) blue GNC. Size distribution images of (g) red GNC, (h) green GNC, and (i) blue GNC. Photographs of as-obtained (j) red GNC, (k) green GNC, and (l) blue GNC. Reproduced with permission from Ref. [14] (Copyright 2015 American Chemical Society). 3.2.3 Sol-gel technique combined with nanoparticle or nanoparticle precursors impregnation Ventura et al. have prepared GNCs made from Au and/or Ag NPs embedded in silica glass matrix [156]. Au and Ag NPs were synthesized by reduction of HAuCl4 and silver nitrate (AgNO3). These NPs were incorporated into the sol solution containing TMOS, which was the precursor to silica glass. The colloidal solutions containing Ag and Au NPs showed a characteristic UV-vis absorption peak at 393 and 520 nm, respectively. Conversely, the colloidal solution with a mixture of NPs exhibited two separate UV-vis absorption peaks corresponding to the absorbance of pure Ag and pure Au NPs. Logically, the absorbance of one component increased as the component concentration increased. The colors of the gels ranged from yellow to 32 Front. Mater. Sci. 2022, 16(3): 220607 deep red with increasing gold concentration. The gels were dried to obtain the colored transparent glasses. AgAu alloys were formed by thermal treatment at temperatures above 200 °C. Both the glasses containing alloys and the glasses containing a mixture of NPs exhibited the orange tones between the ruby red color of Au NPs@silica glass composite and the yellow color of Ag NPs@silica glass composite. Glass composites with pure Ag NPs were bleached at temperatures higher than 300 °C. A small amount of Au NPs was found to avoid the bleaching process [156]. Rouge et al. have embedded zirconia-coated gold NPs into silica glass matrix [256]. The precursors of SiO2, ZrO2, and Au NPs were tetramethyl orthosilicate (TMOS), zirconium(IV) propoxide, and HAuCl4, respectively. Firstly, previously synthesized Au NPs were embedded in a zirconia sol and subsequently incorporated to a TMOS solution. This new sol was gelled and dried at 120 °C and then annealed in air at 850 °C. Finally, the dry gel was densified in air at 1100 °C, which led to the formation of the composite. The size of Au NPs in annealed gel ranged from 8 and 18 nm. In addition, the zirconia phase was dispersed in the silica matrix. In the densified doped glass, the size of Au NPs was between 10 and 30 nm. Particles of ZrO2 appeared in the silica matrix after annealing at 1100 °C. Zirconia was suggested to protect Au NPs during annealing and densification. At the wavelength of 532 nm, β of zirconia-coated Au NPs@silica composite was −1.37 × 10−10 m·W−1 [256]. Sm2O3 NPs have been embedded in the reticular structure of a borosilicate glass by sol-gel route and impregnation [257]. TEOS and trimethyl borate (C3H9BO3) were used as SiO2 and boron precursors, respectively. The gelation of the mixture of precursors (sol) was carried out at 50 °C for 6 h. The gel was aged for 1 week. Then, the gel was dried at 50 °C for 1 additional week. The dry gel was doped with Sm(NO3)3. This doped dry gel was heated in air and N2:H2 (95:5 vol.%) at 450 °C during 5 and 12 h, respectively. Sm(NO3)3 was transformed to Sm2O3 [271]: 2Sm(NO3 )3 → Sm2 O3 + 6NO2 + 1.5O2 (27) Thus, Sm2O3 NPs@borosilicate glass composite was prepared. In brief, the oxidation and reduction treatments were suggested to remove the organic substance, oxidize Sm(NO3)3, and promote the luminescence of Sm2O3. Embedded Sm2O3 behaved like network modifier. It produced NBOs [272]. Therefore, BO3 was transformed to BO4 by taking NBO from Sm2O3, generating the boric anomaly [273]. Similarly, Sm(NO3)3 was also embedded into a sodium borosilicate glass. The composite was synthesized by doping the dry gel with NaNO3 and subsequently with Sm(NO3)3. Then, the doped dry gel was also exposed to air and N2:H2 atmospheres to produce a Sm2O3 NPs@sodium borosilicate glass composite. It is worth mentioning that the glass transition temperature and the melting temperature of the glass are often reduced by adding network modifiers such as the sodium oxide. In addition, network modifier oxides break and open the reticular structure of glasses, which results in loss of connectivity and, therefore, favors doping [273]. In this study, Sm2O3 and Na2O increased the glass transition temperature due to the boric anomaly. The size of the embedded Sm2O3 was 2.7 and 4.1 nm for the Sm2O3 NPs@borosilicate glass and Sm2O3 NPs@sodium borosilicate glass composites, respectively. The optical band gap of Sm2O3 NPs@borosilicate glass and Sm2O3 NPs@sodium borosilicate glass composites was 3.65‒3.86 and 3.39‒3.57 eV, respectively. Therefore, the optical band gap decreased with increasing the amount of Sm2O3 and Na2O. The observed green luminescence (545 nm) under UV light in Sm3+ doped glasses, which was found to be stable for 6 months, has been associated to the 4F3/2 → 6H7/2 electronic transition of Sm3+ ions [257]. A good dispersion of RE NPs in the glass matrix is necessary to increase the PL efficiency. It should be noted that sol-gel methods allow incorporating a higher concentration of reinforcements than conventional methods. Up to 5‒10 wt.% of RE NPs can be embedded into glasses using sol-gel approaches [274‒279]. The aggregation of RE NPs and the devitrification become significant above these concentrations [280]. 3.3 3.3.1 Ion implantation Introduction Ion implantation is an effective tool for introducing reinforcements into the surface layer of a substrate to a depth of several micrometers [281]. Specifically, ion implantation of metals is a well-established technique to embed metal NPs into glass matrices. Table 4 shows examples of GNCs prepared by ion implantation techniques [282‒293]. This method requires lowtemperature processing, controls the concentration of NPs, and overcomes solubility limits of conventional Ag NPs Silica glass Silica glass Silica glass Silica glass Silica glass Silica glass Chalcohalide glass (48.80 wt.% GeS2, 36.04 wt.% Ga2S3, and 15.16 wt.% KBr) Chalcohalide glass (48.80 wt.% GeS2, 36.04 wt.% Ga2S3, and 15.16 wt.% KBr) Chalcohalide glass (48.80 wt.% GeS2, 36.04 wt.% Ga2S3, and 15.16 wt.% KBr) Chalcohalide glass (48.80 wt.% GeS2, 36.04 wt.% Ga2S3, and 15.16 wt.% KBr) Silica glass Au NPs@silica glass Au NPs@silica glass Au NPs@silica glass Au NPs@silica glass Ag NPs@silica glass Ag NPs@silica glass Ag NPs@chalcohalide glass Silica glass Silica glass Silica glas Silica glas Silica glas Silica glas Silica glas Silica glas Silica glas Borate glass (10.66 wt.% B2O3, 67.12 wt.% PbO, 9.62 wt.% GeO2, 10.71 wt.% Bi2O3, and 1.89 wt.% Pr2O3) Ag NPs@silica glass Ag NPs@silica glass Au NPs@silica glass Au NPs@silica glass Au NPs@silica glass Au NPs@silica glass Au NPs@silica glass Au NPs@silica glass Au NPs@silica glass Au NPs@borate glass Ag NPs@silica glass Ag NPs@chalcohalide glass Ag NPs@chalcohalide glass Ag NPs@chalcohalide glass Ag NPs Host matrix Composite (11.68) (2.8) Au NPs (9.98) (9.14) (3.02) (3.18) (2.66) (2.84) (1) (1) (1.2) Au NPs Au NPs Au NPs Au NPs Au NPs Au NPs Au NPs Ag NPs Ag NPs Ag NPs (300) (200) Ag NPs Ag NPs (150) (100) (2.7) (7.7) (12.6) (5.6) (3.6) Size of NPs (min‒max (mean))/nm (2) Ag NPs Ag NPs Au NPs Au NPs Au NPs Au NPs Reinforcement GNCs prepared by ion implantation techniques [282‒293] Table 4 NLO devices: χ(3)(800 nm) = 1.4 × 10−13 esu NLO devices: χ(3)(800 nm) = 3.3 × 10−13 esu NLO devices: χ(3)(800 nm) = 2.8 × 10−13 esu NLO devices: χ(3)(800 nm) = 2.6 × 10−13 esu ‒ ‒ Application Single implantation: ion implantation — Ag ions of 330 keV, ‒ ion fluence (1 × 1016 ion·cm−2); annealing (at 600 °C for 1 h) Single implantation: ion implantation — Ag ions of 1.2 MeV, ion fluence (1 × 1016 ion·cm−2); annealing (at 600 °C for 1 h) Single implantation: ion implantation — Ag ions of 1.7 MeV, ion fluence (1 × 1016 ion·cm−2); annealing (at 600 °C for 1 h) Single implantation: ion implantation — Au ions of 60 keV, Optoelectronics ion fluence (3 × 1016 ion·cm−2), flux (1.5 µA·cm−2) Single implantation: ion implantation — Au ions of 60 keV, ion fluence (4 × 1016 ion·cm−2), flux (1.5 µA·cm−2) Single implantation: ion implantation — Au ions of 60 keV, ion fluence (5 × 1016 ion·cm−2), flux (1.5 µA·cm−2) Single implantation: ion implantation — Au ions of 60 keV, ion fluence (6 × 1016 ion·cm−2), flux (1.5 µA·cm−2) Single implantation: ion implantation — Au ions of 60 keV, ion fluence (7 × 1016 ion·cm−2); flux (1.5 µA·cm−2) Single implantation: ion implantation — Au ions of 60 keV, ion fluence (8 × 1016 ion·cm−2), flux (1.5 µA·cm−2) Single implantation: ion implantation — Au ions of 60 keV, ion fluence (9 × 1016 ion·cm−2), flux (1.5 µA·cm−2) Single implantation: Melt-quenching for glass preparation Gain media for amplifiers in (melting at 1250 °C for 1 h); annealing (at 330 °C for 2 h); ion optical telecommunication implantation — Au ions of 300 keV, ion fluence windows (1 × 1016 ion·cm−2); annealing (at 330 °C for 12 h) Single implantation: ion implantation — Ag ions of 200 MeV, ion fluence (2 × 1017 ion·cm−2) Single implantation: ion implantation — Ag ions of 200 MeV, ion fluence (1 × 1017 ion·cm−2) Single implantation: ion implantation — Ag ions of 200 MeV, ion fluence (5 × 1016 ion·cm−2) Single implantation: ion implantation — Au ions of 190 keV, ion fluence (4 × 1016 ion·cm−2), flux (2 µA·cm−2) Single implantation: ion implantation — Au ions of 190 keV, ion fluence (4 × 1016 ion·cm−2), flux (2 µA·cm−2); annealing (at 900 °C for 1 h) Single implantation: ion implantation — Au ions of 190 keV, 16 ion fluence (4 × 10 ion·cm−2), flux (2 µA·cm−2); annealing (at 900 °C for 3 h) Single implantation: ion implantation — Au ions of 190 keV, 16 ion fluence (4 × 10 ion·cm−2), flux (2 µA·cm−2); annealing (at 900 °C for 12 h) Single implantation: ion implantation — Ag ions of 1 MeV, ion fluence (5 × 1016 ion·cm−2), flux (1 µA·cm−2) Single implantation: ion implantation — Ag ions of 1 MeV, ion fluence (5 × 1016 ion·cm−2), flux (1 µA·cm−2); ion implantation — Si ions of 9.4 MeV, ion fluence (1 × 1014‒8 × 1015 ion·cm−2) Single implantation: ion implantation — Ag ions of 200 MeV, ion fluence (1 × 1016 ion·cm−2) Ion implantation technique [287] [286] [285] [284] [283] [282] Ref. Javier FONSECA. NPs embedded into glass matrices: GNCs 33 Host matrix Silica glass Silica glass Silica glass Silica glass Silica glass Silica glass Silica glass Silica glass Silica glass Composite Cu NPs@silica glass Cu NPs@silica glass Au NPs@silica glass Cu NPs@silica glass Ag NPs@silica glass Ag NPs/Ni NPs@silica glass Ag NPs@silica glass Ag NPs@silica glass Ag NPs/Zn NPs/AgZn NPs@silica glass Ag NPs, Zn NPs and AgZn alloy NPs Ag NPs Ag NPs Ag NPs and Ni NPs Ag NPs Cu NPs Au NPs Cu NPs Cu NPs Reinforcement 2‒12 4‒8 3‒17 0.5‒2 nm Ag NPs, 0.5‒4 nm Ni NPs 35‒48 1‒15 (5.6) 6‒10 6‒10 Size of NPs (min‒max (mean))/nm (20) Application Single implantation: ion implantation — Cu ions of 40 keV, ‒ ion fluence (5 × 1016 ion·cm−2), flux (5 µA·cm−2) Single implantation: ion implantation — Cu ions of 200 keV, NLO devices: χ(3)(532 nm) = ion fluence (3 × 1016 ion·cm−2); annealing (at 400 °C for 1 h) 1.5 × 10−8 esu Single implantation: ion implantation — Au ions of 1.5 MeV, NLO devices: χ(3)(532 nm) = ion fluence (1 × 1017 ion·cm−2); annealing (at 400 °C for 1 h) −3.7 × 10−12 esu Single implantation: ion implantation — Cu ions of 180 keV, NLO devices: χ(3)(532 nm) = ion fluence (1 × 1017 ion·cm−2), flux (1.5 µA·cm−2) −2.1 × 10−7 esu, χ(3)(1064 nm) = −1.2 × 10−7 esu Multiple implantation: ion implantation — Ag ions of ‒ 200 keV, ion fluence (5 × 1016 ion·cm−2), flux (1 µA·cm−2); ion implantation — Cu ions of 110 keV, ion fluence (5 × 1016 ion·cm−2), flux (1.5 µA·cm−2) Multiple implantation:ion on implantation — Ni ions of Optoelectronics, optical 60 keV, ion fluence (5 × 1016 ion·cm−2), flux (4 µA·cm−2); sensors, and antibacterial ion implantation — Ag ions of 70 keV, ion fluence materials (5 × 1016 ion·cm−2), flux (4 µA·cm−2); annealing (at 600 °C for 1 h) Single implantation: ion implantation — Ag ions of 35 keV, Optoelectronics, optical ion fluence (5 × 1016 ion·cm−2), flux (4 µA·cm−2) sensors, and antibacterial materials Multiple implantation: ion implantation — Zn ions of 50 keV, ion fluence (5 × 1016 ion·cm−2), flux (4 µA·cm−2); ion implantation — Ag ions of 35 keV, ion fluence (5 × 1016 ion·cm−2), flux (4 µA·cm−2) Multiple implantation: ion implantation — Ag ions of 35 keV, ion fluence (5 × 1016 ion·cm−2), flux (4 µA·cm−2); ion implantation — Zn ions of 50 keV, ion fluence (5 × 1016 ion·cm−2), flux (4 µA·cm−2) Ion implantation technique [293] [292] [291] [290] [289] [288] Ref. (continued) 34 Front. Mater. Sci. 2022, 16(3): 220607 Javier FONSECA. NPs embedded into glass matrices: GNCs techniques [163]. Ion implantation reaches a high metal filling (beyond the equilibrium limit of metal solubility) in an irradiated glass matrix. It should be noted that high filling factors favor a high χ(3), when the other GNC parameters are maintained. However, ion implantation does not allow an effective control on the size distribution of NPs within the matrix. In addition, this technique does not enable uniform depth of penetration of implanted ions into the matrix [281,294]. Post-implantation thermal or laser annealing are often applied to control the size distribution and penetration of NPs embedded in glass matrices. Embedded metal ions tend to form neutrally charged metal atoms. Metal ions can be neutralized by forming chemical bonds with radicals and ions of the glass or by participating in oxidation processes. Considering the difference in the Gibbs free energies, metal-metal bonds are predominantly created. In other words, clusters of metal atoms are mainly formed. These clusters serve as nuclei of metal NPs. The growth of the metal NPs depends on both the diffusion coefficient and the local concentration of metal atoms. Since diffusion can occur through defects, their production under irradiation is crucial for the formation of NPs. Under diffusion-limited conditions, NPs grow predominantly by joining new implanted metal ions. Moreover, under these conditions, large NPs grow while smaller NPs dissolve. This growth mechanism is known as coarsening and is governed by the Gibbs‒Thomson effect, where the chemical potential is inversely proportional to the particle size. When growth occurs under diffusion-limited and mass conservation conditions, the coarsening mechanism is called Ostwald ripening [282]. The energy of the irradiation is transferred to the glass matrix by electron shell excitation (ionization) and nuclear collisions. Therefore, ion implantation may induce modifications of the glass matrix such as extended and point defects, amorphization and local crystallization, precipitation of a new phase made up of host atoms or implanted ions, etc. The degree of glass modification induced by ion implantation depends on the chemical and structural composition of the glass, on variations in implantation parameters such as type, fluence and energy of ion species, and on process temperature [281]. The implant amount is determined by ion dose (F0). The ions implanted through low-dose irradiation (F0 ≤ 5.0 × 1014 ions·cm−2) are usually well-dispersed 35 throughout the glass matrix and well separated from each other. These ions remain isolated within the glass matrix. When the irradiation ranges from 1015 to 1017 ions·cm−2, the implant concentration often exceeds the solubility limit of metal atoms in glass matrix. Consequently, the reinforcements tend to relax by nucleation and growth. This causes the formation of isolated NPs. At high-dose irradiation (F0 ≥ 1017 ions·cm−2), metal NPs coalescence and form either aggregates or thin quasi-continuous films at the glass surface [281]. The size of metal NPs formed at different depths is proportional to the metal filling factor at the specific depth, which, in turn, depends on the ion irradiation. Generally, different concentrations of embedded atoms are found from layer to layer in an irradiated glass matrix. Large NPs are usually close to the surface of the glass matrix, while subnanometric metal clusters and small NPs are found within the glass matrix but far from the irradiated surface. When applying high dose ion irradiation, the largest NPs are found on the surface due to surface sputtering (Fig. 18) [295]. This size distribution of NPs has a decisive effect on the properties of the GNC. Therefore, to control the properties of GNCs for specific application purposes, the material designer must carefully manipulate the synthesis of such GNCs by ion irradiation. The fluence-dependent depth distribution function (G(z)) for implanted species can be expressed according to Eq. (28) if the implantation probability function is Gaussian-like [296]: N G(z) = 2Y z − Rp + ΦY/ N z − Rp erf √ − erf √ (28) 2 · ∆Rp 2 · ∆Rp where z is the depth coordinate with respect to the surface, N is the atomic density of the substrate, Y is the sputtering yield, Φ is the ion fluence, Rp is the projected range of ions, and ΔRp is the range straggling related to Rp. 3.3.2 Single implantation In an early study, Stepanov et al. embedded Ag NPs in silica glass by ion implantation [295]. 30 keV Ag+ ions with a dose of 5×1016 ions·cm−2 were implanted in silica glass at ion current densities ranging from 4 to 15 µA·cm−2. The size of the embedded NPs within the glass matrix was found to increase as the ion current density increased. An increase in the ion current density caused a temperature rise in the glass matrix that elevated the 36 Front. Mater. Sci. 2022, 16(3): 220607 Fig. 18 Physical phenomena involved in the formation of metal NPs by ion implantation. Reproduced with permission from Ref. [295] (Copyright 2004 Elsevier). mobility of the atoms, thus favoring the growth of NPs [295]. Srivastava et al. have also incorporated Ag NPs into a silica glass by ion implantation of Ag+ ions [283]. Ag+ ions of 1 MeV were implanted in the silica matrix at a fluence of 5 × 1016 ions·cm−2. The Ag+ current density was kept around 1 µA·cm−2. Ag NPs with an average size of 7.7 nm were formed within the silica glass (Fig. 19(a)). The as-implanted composite was bombarded with Si5+ ions of 9.4 MeV at fluences ranging from 1 × 1014 to 8 × 1015 ions·cm−2. Ag NPs were found to dissolve by irradiation of Si ions. The average size of the Ag NPs reduced to 2.7 nm (Fig. 19(b)). The UV-vis optical absorption of the GNC was found to change drastically when the material was irradiated with high-energy Si ions [283]. Metal ions have also been implanted in glasses other than silica glass. Liu et al. have embedded Ag NPs in a chalcohalide glass (GeS2‒Ga2S3‒KBr) by ion implantation [284]. This glass, which was prepared by meltquenching method, was bombarded with Ag+ ions of 200 keV at various fluences (1 × 1016, 5 × 1016, 1×1017, and 2 × 1017 ions·cm−2), thus forming Ag NPs within the glass matrix. The average size of the Ag NPs was 100, 150, 200, and 300 nm within the GNCs prepared with a fluence of 1 × 1016, 5 × 1016, 1 × 1017, and 2 × 1017 ions·cm−2, respectively. Therefore, the size of the NPs increased as the fluence increased. Furthermore, χ(3) of the GNCs improved as the Ag NPs embedded became larger. Thus, at 800 nm, χ(3) values were reported to be 2.6 × 10−13, 2.8 × 10−13, 3.3 × 10−13, and 1.4 × 10−13 esu for the GNCs prepared with fluence values of 1 × 1016, 5 × 1016, 1 × 1017, and 2 × 1017 ions·cm−2, respectively. χ(3) of the GNC prepared with a fluence of 2 × 1017 ions·cm−2 was lower than that of the GNC prepared with a fluence of 1 × 1017 ions·cm−2 due to the greater Fig. 19 TEM images of Ag NPs within silica glass (a) before and (b) after Si ion-irradiation. NP size distribution histograms are shown as insets. Reproduced with permission from Ref. [283] (Copyright 2014 Elsevier). interaction of Ag NPs in the first mentioned GNC [284]. The effect of ion implantation energy on the nucleation of Ag NPs within different glasses (silica glass, BK7 (SiO2‒B2O3‒Na2O‒K2O‒As2O3‒BaO), GIL49 (SiO2‒ Na2O‒K2O‒Al2O3‒CaO‒MgO), and Glass B (SiO2‒Na2O‒ Al2O3)) have been explored [285]. These glasses were implanted with Ag+ ions of 0.33, 1.2, and 1.7 MeV at a constant fluence of 1 × 1016 ions·cm−2. The as-implanted GNCs were annealed at 600 °C for 1 h in air. In silica glass, the average size of the embedded Ag NPs was approximately 1 nm for 1.2 and 1.7 MeV Ag+ implantations. The average size of the Ag NPs within the same glass was around 1.2 nm for 330 keV Ag+ implantation. In those glasses containing Na+ (BK7, GIL49 and Glass B), Ag NPs were prepared at Ag+ energies of 1.2 and 1.7 MeV. However, the subsequent annealing of these GNCs dissolved the NPs already created. In other words, no Ag NPs were found after the annealing procedure in glasses containing Na+ ions. This was explained by an easy migration of Ag+ ions due to Javier FONSECA. NPs embedded into glass matrices: GNCs Ag+‒Na+ ion-exchange in these glasses. The location of the Ag NPs was found to be deeper within the glasses as the ion implantation energy increased. In addition, the location of the Ag NPs was also found to depend on the density of the glasses. The location of the Ag NPs was deeper within a glass as the density of the glasses decreased [285]. In a very similar study, Stanek et al. have embedded Ag NPs in Er3+-doped silicate glasses (SiO2‒Na2O‒ZnO‒Al2O3) by Ag+ ion implantation and subsequent annealing [297]. The glasses with different Er3+ content (0.1, 0.2, and 0.25 mol.% of Er) were previously prepared by melt-quenching. The glasses were bombarded with Ag+ ions of 1.2 and 1.7 MeV at the fluence of 1 × 1016 ions·cm−2. In doing so, Ag NPs with a size of approximately 1‒2 nm were embedded within the glasses. The GNCs were annealed at 600 °C for 1 h in an air atmosphere. This post-implantation annealing dissolved the Ag NPs. It should be noted that regardless of the ion implantation energy, the as-implanted samples exhibited the PL intensity almost three times that of nonimplanted glasses. The PL intensity remained the same in the annealed samples although the Ag NPs were no longer present [297]. Au NPs have also been incorporated into silica glass by ion implantation [286]. Au ions of 60 keV were implanted into the glass matrix. The ion fluence was varied from 3 × 1016 to 9 × 1016 ions·cm−2 with a flux of 1.5 µA·cm−2. The Au NP size was found to be control by such variation of the total fluence. The average size was 2.84, 2.66, 3.18, 37 and 3.02 nm for an ion fluence of 3 × 1016, 4 × 1016, 5 × 1016, and 6 × 1016 ions·cm−2, respectively. The amount of Au NPs increased as ion fluence increased, but without an increase in size. However, the average size of the Au NPs was 9.14, 9.98, and 11.68 nm for an ion fluence of 7 × 1016, 8 × 1016, and 9 × 1016 ions·cm−2, respectively. Therefore, the NPs grew with these higher implantation fluences. It is worth noting that only a part of the NPs grew, while the rest maintained their size. In other words, for these three ion fluences, the larger particles were accompanied by smaller ones. The Ostwald ripening mechanism was suggested in view of this inhomogeneous size distribution [298]. The Au NPs were found to be located near the upper surface of the matrix. In this study, the NLO properties of the GNC prepared with an ion fluence of 6 × 1016 ions·cm−2 were also explored [286]. Marchi et al. have investigated the growth kinetics of Au NPs embedded in silica glass [282]. The GNC (AuNPs/silica glass) was prepared by Au+ ion implantation. 190 keV Au+ ions with a fluence of 4 × 1016 ions·cm−2 and a current density of less than 2 µA·cm−2 were implanted in silica glass. Subsequently, the as-implanted GNC was annealed in an air atmosphere at 900 °C for different times (1, 3, and 12 h). The mean size of the as-implanted GNC and the GNCs annealed during 1, 3, and 12 h was 2, 3.6, 5.6, and 12.6 nm, respectively. Therefore, the mean size of the NPs increased as the annealing time increased (Fig. 20). The Fig. 20 Cross-sectional bright-field TEM images of (a) the as-implanted and the annealed for (c) 1 h, (e) 3 h, and (g) 12 h GNCs. Au NPs size distribution in (b) the as-implanted and the annealed for (d) 1 h, (f) 3 h, and (h) 12 h GNCs. Reproduced with permission from Ref. [282] (Copyright 2002 AIP Publishing). 38 Front. Mater. Sci. 2022, 16(3): 220607 Au NPs in the GNCs were present at a depth of approximately 120 nm below the surface. The growth kinetics of NPs was found to change with annealing times. NPs grew mainly following the diffusion limited mechanism when the heating time was less than 4 h. However, when the annealing time exceeded 5 h, the NPs grew mainly following the Ostwald ripening mechanism. Therefore, there was a transition from a diffusion-limited mechanism to the Ostwald ripening mechanism in the growth kinetics of the Au NPs during annealing time between 4 and 5 h [282]. Au NPs have been implanted into a Pr3+-doped B2O3‒PbO‒Bi2O3‒GeO2 glass [287]. The Pr3+-doped glass was prepared by melt-quenching technique. Au+ ions of 300 keV were implanted into the glass matrix. The ion fluence was 1 × 1016 ions·cm−2. This composite was annealed at 330 °C for 4 and 12 h to form the Au NPs. The average size of Au NPs was 2.8 nm in the GNC obtained after annealing for 12 h. The annealing was found to crystallize a nanometer thick layer of Pb3Ge2O7, Bi4Ge3O12, and Bi2GeO5 at the composite surface due to the defects created by ion implantation. It should be mentioned that defects created during ion implantation can act as nucleation centers promoting crystallization under heat treatment [299]. The luminescent properties of this GNC were also investigated in this study [287]. Beyond Ag and Au NPs, many other metal NPs have been incorporated into glass matrices by ion implantation [288,300]. For instance, the implantation of Cu+ ions within silica glass has been extensively studied [289‒290]. Ghosh et al. have incorporated Cu NPs in silica glass by Cu+ ion implantation [289]. Cu+ ions of 200 keV at a dose of 3 × 1016 ions·cm−2 were implanted in the glass. The as-implanted glass was subsequently annealed in Ar atmosphere at 400 °C for 1 h, thus generating Cu NPs within the glass matrix. In the same study, Au NPs were also obtained within a silica glass by 1.5 MeV Au+ ion implantation with a dose of 1 × 1017 ions·cm−2 and subsequent annealing in Ar atmosphere at 400 °C for 1 h. The size of the NPs prepared in both composites (Cu NPs@silica glass and Au NPs@silica glass) was found to range between 6 and 10 nm. At the wavelength of 532 nm, the χ(3) was 1.456 × 10−8 and 3.66 × 10−12 esu at 532.8 nm for Cu NPs@silica glass and Au NPs@silica glass, respectively [289]. Similarly, Wang et al. have prepared Cu NPs in silica glass by implantation of Cu+ ions [290]. The glass matrix was bombarded with 180 keV Cu+ ions with a fluence of 1×1017 ions·cm−2 and a current density of 1.5 µA·cm−2. Thus, Cu NPs with a size ranging from 1 to 15 nm and an average size of 5.6 nm were formed within the glass matrix. The χ(3) of this GNC was −2.1 × 10−7 and −1.2 × 10−7 esu at 532 and 1064 nm, respectively [290]. 3.3.3 Multiple implantation The preparation of GNCs with different metals or metal alloys through multiple implantation experiments has attracted much attention. In an early study, Gonella et al. have prepared AuCu alloy NPs embedded in silica glass by dual ion implantation and heat treatment [301]. Au+ ions of 190 keV (3 × 1016 ions·cm−2) and subsequently Cu+ ions of 90 keV (3 × 1016 ions·cm−2) were implanted in silica glass. Current densities were kept below 2 µA·cm−2. Thus, AuCu alloy NPs were formed in the asimplanted GNC. Larger AuCu alloy NPs were found when the as-implanted GNC was annealed in an H2 atmosphere at 900 °C for 1 h. The as-implanted GNC was also annealed in an air atmosphere at 900 °C for 1 h. In doing so, Cu migrated to the surface of the glass matrix where it was oxidized to form CuO and Cu2O, while Au+ ions diffused deeper into the matrix [301]. Therefore, the GNC was modified by selective annealing in a reducing or oxidizing atmosphere. In another study, AuAg alloy NPs and nanoplanets (NPLs) have been embedded in silica glass by dual ion implantation [302]. It should be mentioned that NPLs consist of a central NP surrounded by small satellite NPs [303]. 190 keV Au+ ions were implanted into the glass matrix with a fluence 3 × 1016 ions·cm−2. The current density was 2 µA·cm−2. Then, 130 keV Ag+ ions were implanted into the matrix with a fluence of 3 × 1016 ions·cm−2. The current density was kept at 2 µA·cm−2. The as-implanted sample was annealed at 800 °C for 1 h, forming AuAg alloy NPs with an average size of 12 nm. This GNC (AuAg alloy NPs@silica glass) was irradiated with Ar at an energy of 190 keV, a current density of 0.2 µA·cm−2 and a fluence of 2.5 × 1016 ions·cm−2. In doing so, NPs became NPLs. Au-rich satellite NPs were formed with a mean size of 2 nm at 2‒3 nm from the original NPs, thus developing a NPL structure [302]. Multiple implantations can also be employed to synthesize monometallic NPs without any further treatment. These monometallic NPs can be quite distinctive in shape, size, and spatial distributions due to the influence of the previous or subsequent implantation Javier FONSECA. NPs embedded into glass matrices: GNCs of the other ions. For example, Xiao et al. have incorporated Ag NPs into a silica glass by Ag+ ion implantation and Cu+ ion post-implantation [291]. The implantation fluences of Ag+ and Cu+ ions were 5 × 1016 ions·cm−2. The flux density of Ag was 1 µA·cm−2 with an energy of 200 keV. The flux density of Cu was 1.5 µA·cm−2 with an energy of 110 keV. Spherical Ag NPs with a size ranging from 35 to 48 nm were found to be aligned at the same depth in the silica glass. Cu NPs were not detected. However, they may have existed but with a much smaller size than the Ag NPs. The Ag NPs formed were bombarded by Cu+ ions that induced thermal diffusion, which caused Ag+ to exceed the solubility limit and form large Ag NP. The thermal diffusion was also suggested to increase the diffusion capacity of Cu ions. Therefore, dual ions implantation offered some control of the size and position of the NP in layered structures. In contrast, a wide Ag NP size distribution was found when silica glass was only bombarded with Ag ions (200 keV at fluences of 5 × 1016 and 1.0 × 1017 ions·cm−2, and a flux of 1 µA·cm−2). Large NPs were found near the surface and small NPs were found deep within the glass matrix [291]. This distribution of NPs may be explained by the following guidelines: (1) Small NPs near the surface of the glass matrix are sputtered by the following implanted ions. It should be remembered that the sputtered effect is proportional to fluence. (2) Defects in the glass matrix occur during implantation. These defects are close to the surface of the matrix. Implanted ions are easily trapped in these defects to form numerous metal nuclei. Therefore, larger NPs are formed near the surface of the matrix. (3) Metal ions penetrate deep into the matrix due to sputtering. Silica glass was implanted with Ni2+ ions of 60 keV at fluences of 1 × 1016, 5 × 1016, and 1 × 1017 ions·cm−2 and then with Ag+ ions of 70 keV at the fluence of 5.0 × 1016 ions·cm−2 [292]. The implantation was carried out at a beam current density of 4 µA·cm−2. Subsequently, the composites were annealed at 600 °C for 1 h in Ar atmosphere. The pre-implanted Ni2+ ions produced nucleation sites that affected post-implanted Ag+ ions, Ag nucleation, and NP growth. Specifically, the preimplantation of Ni improved the migration and nucleation process of the Ag implants. Furthermore, this preimplantation decreased the sputtering rate of Ag postimplantation. At a low fluence of 1.0 × 1016 Ni2+ 39 ions·cm−2, a small number of Ni2+ ions were implanted and therefore the quantity of defects produced to provide nucleation sites was low. After the annealing at 600 °C, the Ni nuclei formed were tiny and highly dispersed. Furthermore, Ag NPs with a wide size distribution appeared on the surface of the glass matrix. The concentration of Ag increased when applying a preimplantation of Ni2+ ions at a fluence of 5.0 × 1016 ions·cm−2. This GNC exhibited the highest Ag concentration. During Ag+ ion implantation, Ni NPs near the surface of the composite were removed by sputtering. In this GNC, while the size of Ni NPs ranged between 0.5 and 4 nm, the size of the Ag NPs ranged between 0.5 and 2 nm. It was also suggested that AgNi alloy NPs were formed. At the high fluence of 1.0 × 1017 Ni2+ ions·cm−2, many Ni0 atoms were implanted near the glass surface, which trapped Ag0 atoms during the Ag postimplantation. The Ag0 atoms were trapped near the surface, favoring heavy sputtering effects [292]. It is worth noting that the sputtered thickness is often proportional to the ion fluence [304]. To synthesize this composite without annealing, Yamada et al. have performed a dual implantation with 380 keV Ag+ ions and then with 200 keV Ni2+ ions on silica glass [305]. The ion fluences were 3 × 1016 and 6 × 1016 ions·cm−2 for Ag+ and Ni2+ ions, respectively. Ag NPs were found within the glass after Ag+ ion implantation. Then Ni2+ ions were implanted at almost the same depth. Ni2+ ion implantation was found to affect the SPR peak of Ag NPs. The authors suggested three possible reasons for the change in Ag NPs after Ni2+ ion implantation: (1) synthesis of composite NPs that include not only Ag0 atoms but also Ni0 atoms; (2) change in Ag NPs morphology; and/or (3) change in n2 of the glass matrix due to implantation-induced lattice defects. Liu’s research group has prepared different GNCs by dual implantation of Zn2+ and Ag+ ions in different sequences (Fig. 21) [293]. Spherical and ellipsoidal Ag NPs were embedded in silica glass by Ag+ ion implantation (Ag ions of 35 keV; 5 × 1016 Ag ions·cm−2; 4 µA·cm−2). The layer containing NPs was about 40 nm thick. The size of the Ag NPs ranged between 3 and 17 nm. Large NPs were found near the surface, while small NPs were found in a deeper region. When the silica glass was bombarded with Zn2+ ions and then Ag+ ions (Zn2+ ions of 50 keV; 5 × 1016 Zn2+ ions·cm−2; Ag+ ions of 35 keV; 5 × 1016 Ag+ ions·cm−2; 4 µA·cm−2), spherical NPs with a size ranging from 4 to 8 nm were formed in a 40 Front. Mater. Sci. 2022, 16(3): 220607 Fig. 21 (a) Cross-sectional TEM image, (b) SAED result, and (c) Ag NP size distribution of the GNC prepared by Ag+ ion implantation. (d) Cross-sectional TEM image, (e) SAED result, and (f) Ag NP size distribution of the GNC prepared by Zn2+ ion implantation and Ag+ ion post-implantation. (g) Cross-sectional TEM image, (h) SAED result, and (i) Ag NP size distribution of the GNC prepared by Ag+ ion implantation and Zn2+ ion post-implantation. Reproduced with permission from Ref. [293] (Copyright 2013 AIP Publishing). 35 nm thick region. Most NPs were reported to be metallic Ag. Therefore, this dual implantation provided a smaller NP size and narrower size distribution. The Zn2+ ion implantation formed Zn-related compounds in the matrix and also generated defects. These products and defects favored the subsequent nucleation of Ag and promoted the diffusion of Ag0 atoms [306‒307]. An inverse implantation sequence (first Ag+ ion implantation and then Zn2+ ion implantation) generated NPs from 2 to 12 nm. The layer containing NPs was about 53 nm thick. These NPs were suggested to be Ag NPs, Zn NPs, and AgZn alloy NPs [293]. This group has also explored the effects of Zn2+ ion post-implantation on Cu NPs embedded in silica glass [308‒309]. Cu NPs were incorporated into silica glass by 45 keV Cu+ ion implantation at a fluence of 1.0 × 1017 ions·cm−2 and a flux density of 4 µA·cm−2. Spherical Cu NPs were formed and distributed in a depth range of 4 to 55 nm. The Cu NPs exhibited a size distribution of 2 to 17 nm with a mean size of (6.9 ± 3.1) nm. This GNC was subsequently bombarded with 50 keV Zn2+ ions at fluences of 1 × 1016, 5 × 1016, and 1 × 1017 ions·cm−2. The flux density was kept at 4 µA·cm−2. At the low fluence of 1.0 × 1016 ions·cm−2, the loss of Cu by sputtering was negligible. The NPs were still mainly Cu NPs. At the fluence of 1.0 × 1017 ions·cm−2, most of the pre-implanted Cu atoms were sputtered. It resulted in the formation of Zn NPs. At the intermediate fluence of 5.0 × 1016 ions·cm−2, the sputtering effect induced a significant loss of Cu. Large Cu NPs were found near the surface of the glass matrix and CuZn alloy NPs were found in the depth range of about 7‒28 nm. In the deepest region, there were small Zn NPs and/or Zn-rich alloy NPs. The GNC prepared by applying a post-implantation with intermediate fluence (5.0 × 1016 ions·cm−2) was annealed in a N2 atmosphere. Zn NPs grew by Ostwald ripening during low temperature annealing [310]. Zn-rich CuZn alloy NPs were formed by annealing at temperatures of 300 to 350 °C (Fig. 22(b)). There was an increase in the Zn content in the CuZn alloy NPs due to a thermally driven diffusion of Zn0 atoms toward these NPs. This process was referred to as realloying of the NPs. Core‒shell (Cu-rich core with a Znrich shell) NPs were formed at annealing temperatures in the range of 400 to 450 °C (Fig. 22(c)). This thermal treatment decomposed the CuZn alloy NPs. Moreover, the Zn0 atoms diffused towards the surface of the CuZn alloy NPs according to the Kirkendall effect [311‒313]. This process was called dealloying of the NPs. The decomposition of the CuZn alloy NPs due to Zn diffusion was completed at 500 °C. Thus, the Cu0 atoms formed a core of the new NPs, while the shell of these NPs was made of Zn or Zn-related compounds (Fig. 22(d)) [308,314‒315]. In a similar study, Au core‒Ag shell bimetallic NPs have been prepared into a silica glass by dual Ag+ and Au+ ions implantation and subsequent annealing [316]. Silica glass was bombarded with 1.8 MeV Ag+ ions and 3.7 MeV Au+ ions at the same fluence of 5 × 1016 ions·cm−2. There was a mixture of Javier FONSECA. NPs embedded into glass matrices: GNCs 41 Fig. 22 Schematic illustration of the annealing processes in a GNC formed by Cu+ ion implantation and Zn2+ ion post-implantation (fluence of 5 × 1016 ions·cm−2): (a) CuZn alloy NP; (b) Zn diffusion-driven realloying process; (c) Zn diffusion-driven formation of the Cu-rich core and Zn-rich shell structure; (d) Cu core and Zn shell NP. Reproduced with permission from Ref. [308] (Copyright 2013 American Chemical Society). small Ag NPs and Au atoms in the glass matrix prior to any annealing. The as-prepared GNC were annealed at 900 °C for 1 h in air. In doing so, monometallic Au NPs and Au‒Ag core‒shell NPs with homogeneous distribution appeared in the GNC. The Au NPs acted as nucleation centers. Thus, Ag atoms and NPs formed a continuous shell around the Au NPs through an Ostwald ripening mechanism [316]. In brief, multiple implantations can be used to fabricate core‒shell bimetallic NPs. The sequential ion implantation adds a further degree of freedom to the engineering of new materials. In addition, nonmetals can also be implanted into glasses. Torres‒Torres et al. have incorporated Si QDs and Au NPs into silica glass by ion implantation [317]. The glass was exposed to Si2+ ions of 1.5 MeV at the fluence of 2.5 × 1017 ions·cm−2. Subsequently, the asimplanted glass was annealed in a reducing atmosphere at 1100 °C for 1.5 h. This heat treatment promoted the nucleation and growth of Si QDs within the glass matrix. Au2+ ions of 1.5 MeV at the fluence of 8.5 × 1016 ions·cm−2 were implanted in this GNC (Si QDs@silica glass). A new annealing was carried out in an oxidizing atmosphere at 1000 °C for 1 h to precipitate the Au NPs. Si2+ and Au2+ ions were found to reach 2 and 0.66 µm depth in the glass matrix, respectively. In other words, both ion distributions did not overlap. Therefore, it was shown that it is possible to control the position of the ion distributions by adjusting the energy and the dose of the ions to be implanted [317]. 3.4 Ion-exchange 3.4.1 Introduction The ion-exchange process has been widely used to dope glasses with metal ions and, after promoting the precipitation of these ions, prepare GNCs [157,318‒319]. Ion-exchange is a simple and reproducible technique that allows incorporating up to 1021 atoms·cm−3 in a glass matrix, co-doping the glass matrix with different metal species, and space-selectively precipitating NPs within the glass matrix. Additionally, the ion-exchange process is inexpensive in terms of materials, equipment, and manufacturing process. This technology depends on the ion-exchange parameters such as temperature, glass composition and molten salt solution, and on the post ionexchange treatments such as thermal annealing and laser irradiation techniques. Consequently, these parameters and treatments have been and still are intensively studied to achieve control of the preparation of GNCs through ion-exchange technology. Therefore, ion-exchange is a promising technique for NP formation within a glass matrix. Unfortunately, this process is limited to the incorporation of some metal species and alkali-containing glasses [163]. Ion-exchange is typically carried out by replacing monovalent cations, such as alkali cations (Li+, Na+, K+, etc.), present in a glass with other monovalent cations from a molten salt solution (Fig. 23). The overall chemical reaction of the ion-exchange process is [320]: A+B ↔ A+B (29) where A is the monovalent cation initially contained in the molten salt solution and B is the monovalent cation initially located in the glass matrix. When ion-exchange occurs, A and B become A and B, respectively. Therefore, A and B are the monovalent cations situated in the glass matrix and in the molten salt solution, respectively. In other words, the bar denotes the cations in the glass phase. The equilibrium constant (K) can be 42 Front. Mater. Sci. 2022, 16(3): 220607 Fig. 23 (a) Schematic illustration of thermally assisted ion-exchange. (b) The monovalent cation A is introduced into the glass matrix, which contains the monovalent cation B, by an interphase chemical potential gradient. To maintain charge neutrality, the monovalent cation B is released into the molten salt solution. (c) Schematic illustration of electric field-assisted ion-exchange. expressed as follows: K= aA · aB aB · aA (30) where aA , aA , aB, and aB are the thermodynamic activities of cations A and B in the corresponding phases. A high K value indicates a large uptake of A by the glass matrix [321]. Ion-exchange can be limited by (1) the mass transport of cations in the molten salt solution; (2) the kinetics of adsorption and desorption at the interface; and (3) the mass transport of cations in the glass matrix (Fig. 23) [321]. (1) The mass transport of cations in the molten salt solution. The mass transfer of A from the molten salt solution to the glass-solution interface occurs by convection and diffusion. Although convection can be forced, there is a stagnant boundary near the glasssolution interface where convective mixing does not occur. Here, the mass transfer of A to the glass-solution interface processes through diffusion. Conversely, the mass transfer of B from the glass-solution interface to the molten salt solution occurs first by diffusion through the stagnant boundary and then by convection through the molten salt solution. The parameter α informs about the ease with which diffusion happens: 1/2 N A DA α= NA DA where NA and N A are the mole fraction of A in the molten salt solution and in the glass matrix, respectively. DA and DA are the diffusion coefficient of A in the molten salt solution and in the glass matrix, respectively. When α ≥ 10, the ion-exchange is not limited by mass transfer due to diffusion in the molten salt solution [322]. (2) The kinetics of adsorption and desorption at the interface. Surface kinetics are not likely to be a limiting factor in the ion-exchange process. This action is much faster than mass transport processes in molten salt solution and glass matrix. (3) The mass transport of cations in the glass matrix. In the glass matrix, the mass transport of cations is carried out entirely by diffusion. Therefore, this is a relatively slow action. The cation diffusion depends on the surface boundary condition (Eq. (29)) and the diffusion properties in the glass. According to the Regular Solution Theory [323], the relationship between NA and N A can be expressed as follows [320]: ) N A NA H − ln K (1 − 2NA ) = n ln ln + 1 − NA RT 1 − NA ( where n is given by n= (31) (32) ∂ (ln aA ) lnC A (33) H is the interaction energy of the two cations in the Javier FONSECA. NPs embedded into glass matrices: GNCs 43 molten salt solution, R is the gas constant, T is the temperature, and C A is the concentration of A in the glass matrix. It should be mentioned that n = 1 when the glass has an ideal behavior. As mentioned above, a high K value indicates a large uptake of A by the glass matrix. Once the metal ions have been incorporated into the glass matrix through the ion-exchange process, they can be converted into metal NPs by the process itself or by applying thermal annealing or laser irradiation techniques. Table 5 shows examples of GNCs prepared by ionexchange techniques [79,324‒336]. 3.4.2 Direct formation of NPs Ag NPs have been embedded in a borosilicate glass (SiO2‒B2O3‒Na2O‒Al2O3‒NaF) applying only Ag+‒Na+ ion-exchange [324]. The glass was ion-exchanged in a molten bath of AgNO3:NaNO3 with different molar ratios (1:10, 1:50, 1:200, and 1:1000) at 310 °C for 1 h. In the ion-exchange process, smaller Na+ ions were replaced by larger Ag+ ions, leading to the breaking of Si−O bonds and the generation of structural defects and NBO anions [337]. The state of Ag within the glass matrix was ionic Ag+ when the molar ratio of the molten mixture of AgNO3:NaNO3 was 1:1000 (Fig. 24(a)). However, when the AgNO3:NaNO3 molar ratio increased, Ag NPs appeared within the glass (Fig. 24). The size of the Ag NPs increased as the AgNO3:NaNO3 molar ratio increased. The size of Ag NPs within the glass after ionexchange with an AgNO3:NaNO3 molar ratio of 1:10, 1:50, and 1:200 was (6.2 ± 2.5), (4.2 ± 2.2), and (2.0 ± 1.0) nm, respectively. Similarly, Ag NPs were embedded within a Sm3+-doped borosilicate glass (SiO2‒B2O3‒Na2O‒Al2O3‒NaF‒Sm2O3) by Ag+‒Na+ ionexchange. It is worth mentioning that the Ag NPs precipitated within the Sm3+-doped borosilicate glass when the AgNO3:NaNO3 molar ratio was 1:10. Upon excitation at 260 nm, the PL of Sm3+ was enhanced by the energy transfer from Ag+ ions to Sm3+ ions. However, upon excitation at 337 nm, the PL of Sm3+ was enhanced by the energy transfer from Ag NPs to Sm3+ ions [324]. Quaranta et al. have explored in detail the rearrangement of a sodium borosilicate glass after Ag+‒Na+ ionexchange [338]. In their study, the glass was immersed in a molten mixture of AgNO3:NaNO3 with different molar fractions (0.1:99.9, 0.5:99.5, 1:99, 2:98, 20:80, and 40:60 mol.%) at 330 °C for 1 h. There was a significant decrease in Q3 content and an increase in Q2 content Fig. 24 TEM images of GNCs prepared by ion-exchange in a molten bath with the AgNO3:NaNO3 molar ratio of (a) 1:1000, (b) 1:200, (c) 1:50, and (d) 1:10. The insets are the Ag NP size distribution histograms. Reproduced with permission from Ref. [324] (Copyright 2019 John Wiley and Sons). relative to the pristine glass when the Ag content increased. It is worth mentioning that the Q4 correspond to the fully polymerized structural species (SiO2); the Q3 species are those in which the number of NBO per silicon (NBO/Si) is 1 (Si2O5); and the Q2 species are those in which NBO/Si is 2 (SiO3) [339‒340]. Therefore, a progressive depolymerization of the silica network was found as the Ag content increased. The depolymerization was favored by the self-coordination tendency of Ag+ ions and by the release of the compressive stress energy with the breaking of Si−O bonds [341‒342]. Furthermore, a progressive aggregation of Ag NPs was also found as the Ag content increased. The size of the Ag NPs embedded within the glass increased slightly as the molar ratio of the molten mixture for ion-exchange increased. The mean size of the Ag NPs within those glasses that were ionexchanged with 0.1:99.9, 20:80, and 40:60 AgNO3:NaNO3 mol.% was 1.2, 1.5, and 1.5 nm, respectively. The reduction of Ag+ ions during the ionexchange was related to the depolymerization of the glass network. This correlation has been evidenced by Araujo et al. [343]. Furthermore, the presence of SiO2 in phosphate glasses has also been shown to promote the reduction of Ag+ ions due to the increase in the basicity of Ag NPs@silicate glass Ag NPs@silicate glass Ag NPs@silicate glass Ag NPs@silicate glass Ag NPs@silicate glass Ag NPs@silicate glass Ag NPs@silicate glass Ag NPs@silicate glass Ag NPs@silicate glass Ag NPs@silicate glass Ag NPs@silicate glass Ag NPs@silicate glass Ag NPs@silicate glass Ag NPs@borosilicate glass Ag NPs@borosilicate glass Ag NPs Borosilicate glass (52.70 wt.% SiO2, 11.11 wt.% B2O3, 24.71 wt.% Na2O, 8.13 wt.% Al2O3, and 3.35 wt.% NaF) Borosilicate glass (52.70 wt.% SiO2, 11.11 wt.% B2O3, 24.71 wt.% Na2O, 8.13 wt.% Al2O3, and 3.35 wt.% NaF) Borosilicate glass (52.70 wt.% SiO2, 11.11 wt.% B2O3, 24.71 wt.% Na2O, 8.13 wt.% Al2O3, and 3.35 wt.% NaF) Silicate glass (40.45 wt.% SiO2, 15.26 wt.% Al2O3, 13.91 wt.% Na2O, 6.03 wt.% MgO, and 24.35 wt.% ZnO) Silicate glass (40.45 wt.% SiO2, 15.26 wt.% Al2O3, 13.91 wt.% Na2O, 6.03 wt.% MgO, and 24.35 wt.% ZnO) Silicate glass (40.45 wt.% SiO2, 15.26 wt.% Al2O3, 13.91 wt.% Na2O, 6.03 wt.% MgO, and 24.35 wt.% ZnO) Silicate glass (40.45 wt.% SiO2, 15.26 wt.% Al2O3, 13.91 wt.% Na2O, 6.03 wt.% MgO, and 24.35 wt.% ZnO) Silicate glass (40.45 wt.% SiO2, 15.26 wt.% Al2O3, 13.91 wt.% Na2O, 6.03 wt.% MgO, and 24.35 wt.% ZnO) Silicate glass (40.45 wt.% SiO2, 15.26 wt.% Al2O3, 13.91 wt.% Na2O, 6.03 wt.% MgO, and 24.35 wt.% ZnO) Silicate glass (40.45 wt.% SiO2, 15.26 wt.% Al2O3, 13.91 wt.% Na2O, 6.03 wt.% MgO, and 24.35 wt.% ZnO) Silicate glass (40.45 wt.% SiO2, 15.26 wt.% Al2O3, 13.91 wt.% Na2O, 6.03 wt.% MgO, and 24.35 wt.% ZnO) Silicate glass (40.45 wt.% SiO2, 15.26 wt.% Al2O3, 13.91 wt.% Na2O, 6.03 wt.% MgO, and 24.35 wt.% ZnO) Silicate glass (39.41 wt.% SiO2, 14.86 wt.% Al2O3, 13.55 wt.% Na2O, 5.88 wt.% MgO, 23.73 wt.% ZnO, and 2.57 wt.% Eu2O3) Silicate glass (39.41 wt.% SiO2, 14.86 wt.% Al2O3, 13.55 wt.% Na2O, 5.88 wt.% MgO, 23.73 wt.% ZnO, and 2.57 wt.% Eu2O3) Silicate glass (39.41 wt.% SiO2, 14.86 wt.% Al2O3, 13.55 wt.% Na2O, 5.88 wt.% MgO, 23.73 wt.% ZnO, and 2.57 wt.% Eu2O3) Silicate glass (39.41 wt.% SiO2, 14.86 wt.% Al2O3, 13.55 wt.% Na2O, 5.88 wt.% MgO, 23.73 wt.% ZnO, and 2.57 wt.% Eu2O3) Ag NPs Ag NPs Ag NPs Ag NPs Ag NPs Ag NPs Ag NPs Ag NPs (3.6 ± 1.4) (3.4 ± 1.5) (3.3 ± 1.2) (3.0 ± 1.6) (2.4 ± 1.2) (1.8 ± 1.2) (2.5 ± 1.2) (2.0 ± 1.0) (1.6 ± 1.0) (2.3 ± 1.8) Ag NPs Ag NPs (2.2 ± 1.1) (2.0 ± 1.1) (1.5 ± 1) (6.2 ± 2.5) (4.2 ± 2.2) (2.0 ± 1.00) Size of NPs (minmax (mean))/nm Ag NPs Ag NPs Ag NPs Ag NPs Ag NPs Reinforcement Host matrix GNCs prepared by ion-exchange techniques [79,324–336] Ag NPs@borosilicate glass Composite Table 5 Direct formation of NPs: ion-exchange — AgNO3:NaNO3 melt (0.99:99.01 wt.%), 310 °C for 1 h Direct formation of NPs: ion-exchange — AgNO3:NaNO3 melt (3.84:96.16 wt.%), 310 °C for 1 h Direct formation of NPs: ion-exchange — AgNO3:NaNO3 melt (16.66:83.34 wt.%), 310 °C for 1 h Direct formation of NPs: ion-exchange — AgNO3:NaNO3 melt (1.98:98.02 wt.%), 350 °C for 1 h Direct formation of NPs: ion-exchange — AgNO3:NaNO3 melt (13.08:86.92 wt.%), 350 °C for 1 h Direct formation of NPs: ion-exchange — AgNO3:NaNO3 melt (33.32:66.68 wt.%), 350 °C for 1 h Direct formation of NPs: ion-exchange — AgNO3:NaNO3 melt (66.65:33.35 wt.%), 350 °C for 1 h Direct formation of NPs: ion-exchange — AgNO3:NaNO3 melt (13.08:86.92 wt.%), 380 °C for 15 min Direct formation of NPs: ion-exchange — AgNO3:NaNO3 melt (13.08:86.92 wt.%), 380 °C for 1.5 h Direct formation of NPs: ion-exchange — AgNO3:NaNO3 melt (13.08:86.92 wt.%), 380 °C for 4 h Direct formation of NPs: ion-exchange — AgNO3:NaNO3 melt (13.08:86.92 wt.%), 300 °C for 1 h Direct formation of NPs: ion-exchange — AgNO3:NaNO3 melt (13.08:86.92 wt.%), 400 °C for 1 h Direct formation of NPs: ion-exchange — AgNO3:NaNO3 melt (1.98:98.02 wt.%), 380 °C for 1 h Direct formation of NPs: ion-exchange — AgNO3:NaNO3 melt (13.08:86.92 wt.%), 380 °C for 1 h Direct formation of NPs: ion-exchange — AgNO3:NaNO3 melt (33.32:66.68 wt.%), 380 °C for 1 h Direct formation of NPs: ion-exchange — AgNO3:NaNO3 melt (66.65:33.35 wt.%), 380 °C for 1 h Ion-exchange technique Photoluminescence ‒ ‒ Photoluminescence Application [325] [325] [325] [324] Ref. 44 Front. Mater. Sci. 2022, 16(3): 220607 Silicate glass (71.58 wt.% SiO2, 14.36 wt.% Na2O, 6.59 wt.% CaO, 2.67 wt.% MgO, 3.21 wt.% Al2O3, 0.94 wt.% K2O, 0.4 wt.% SO3, and 0.26 wt.% Fe2O3) Silicate glass (71.58 wt.% SiO2, 14.36 wt.% Na2O, 6.59 wt.% CaO, 2.67 wt.% MgO, 3.21 wt.% Al2O3, 0.94 wt.% K2O, 0.4 wt.% SO3, and 0.26 wt.% Fe2O3) Silicate glass (71.58 wt.% SiO2, 14.36 wt.% Na2O, 6.59 wt.% CaO, 2.67 wt.% MgO, 3.21 wt.% Al2O3, 0.94 wt.% K2O, 0.4 wt.% SO3, and 0.26 wt.% Fe2O3) Silicate glass (71.58 wt.% SiO2, 14.36 wt.% Na2O, 6.59 wt.% CaO, 2.67 wt.% MgO, 3.21 wt.% Al2O3, 0.94 wt.% K2O, 0.4 wt.% SO3, and 0.26 wt.% Fe2O3) Silicate glass (71.58 wt.% SiO2, 14.36 wt.% Na2O, 6.59 wt.% CaO, 2.67 wt.% MgO, 3.21 wt.% Al2O3, 0.94 wt.% K2O, 0.4 wt.% SO3, and 0.26 wt.% Fe2O3) Silicate glass (71.58 wt.% SiO2, 14.36 wt.% Na2O, 6.59 wt.% CaO, 2.67 wt.% MgO, 3.21 wt.% Al2O3, 0.94 wt.% K2O, 0.4 wt.% SO3, and 0.26 wt.% Fe2O3) Soda-lime silicate glass (71.63 wt.% SiO2, 2.08 wt.% Al2O3, 0.5 wt.% Fe2O3, 0.19 wt.% TiO2, 0.59 wt.% SO3, 6.27 wt.% CaO, 2.8 wt.% MgO, 15.4 wt.% Na2O, and 0.53 wt.% K2O) Soda-lime silicate glass (71.63 wt.% SiO2, 2.08 wt.% Al2O3, 0.5 wt.% Fe2O3, 0.19 wt.% TiO2, 0.59 wt.% SO3, 6.27 wt.% CaO, 2.8 wt.% MgO, 15.4 wt.% Na2O, and 0.53 wt.% K2O) Soda-lime silicate glass (71.63 wt.% SiO2, 2.08 wt.% Al2O3, 0.5 wt.% Fe2O3, 0.19 wt.% TiO2, 0.59 wt.% SO3, 6.27 wt.% CaO, 2.8 wt.% MgO, 15.4 wt.% Na2O, and 0.53 wt.% K2O) Soda-lime silicate glass (73.65 wt.% SiO2, 14.64 wt.% Na2O, 3.2 wt.% Al2O3, 7.5 wt.% CaO, and 1.0 wt.% MgO) Ag NPs@silicate glass Ag NPs@silicate glass Ag NPs@silicate glass Ag NPs@silicate glass Ag NPs@silicate glass Ag NPs@soda-lime silicate glass Ag NPs@soda-lime silicate glass Ag NPs@soda-lime silicate glass Ag NPs@soda-lime silicate glass Corning 0211 glass Host matrix Ag NPs@silicate glass Ag NPs@Corning 0211 glass Composite Ag NPs Ag NPs Ag NPs Ag NPs Ag NPs Ag NPs Ag NPs Ag NPs Ag NPs Ag NPs Ag NPs Reinforcement (2) (2.2 ± 0.6) (1.9 ± 1.2) (1.8 ± 1.1) (7.2) (4.4) (2.9) (4.0 ± 1.5) (3.3 ± 1.2) (2.8 ± 1) (10) Size of NPs (minmax (mean))/nm Application Direct formation of NPs: ion-exchange NLO devices: β(800 nm) = — AgNO3:NaNO3:KNO3 melt 1.4 × 10−14 cm2·W−1 (8.77:41.66:49.57 wt.%), 300 °C for 6 h; Al film evaporation; immersion in a KNO3 salt at 400 °C for 2 h Thermal annealing in air: ion-exchange Photoluminescence — AgNO3:NaNO3 melt (13.08:86.92 wt.%), 370 °C for few min; post ion-exchange annealing — air atmosphere (450 °C for 1 h) Thermal annealing in air: ion-exchange — AgNO3:NaNO3 melt (13.08:86.92 wt.%), 370 °C for few min; post ion-exchange annealing — air atmosphere (500 °C for 1 h) Thermal annealing in air: ion-exchange — AgNO3:NaNO3 melt (13.08:86.92 wt.%), 370 °C for few min; post ion-exchange annealing — air atmosphere (550 °C for 1 h) Thermal annealing in air: ion-exchange Photoluminescence — AgNO3:NaNO3 melt (13.08:86.92 wt.%), 370 °C for 2 min; post ion-exchange annealing — air atmosphere (500 °C for 1 h) Thermal annealing in air: ion-exchange — AgNO3:NaNO3 melt (13.08:86.92 wt.%), 370 °C for 2 min; post ion-exchange annealing — air atmosphere (550 °C for 1 h) Thermal annealing in air: ion-exchange — AgNO3:NaNO3 melt (13.08:86.92 wt.%), 370 °C for 2 min; post ion-exchange annealing — air atmosphere (600 °C for 1 h) Thermal annealing in air: ion-exchange ‒ — AgNO3:NaNO3 melt (33.33:66.66 wt.%), 360 °C for 1 h; post ion-exchange annealing — air atmosphere (450 °C for 30 min) Thermal annealing in air: ion-exchange — AgNO3:NaNO3 melt (33.33:66.66 wt.%), 360 °C for 1 h; post ion-exchange annealing — air atmosphere (500 °C for 30 min) Thermal annealing in air: ion-exchange — AgNO3:NaNO3 melt (33.33:66.66 wt.%), 360 °C for 1 h; post ion-exchange annealing — air atmosphere (550 °C for 30 min) Thermal annealing in air: ion-exchange ‒ — AgNO3:NaNO3 melt (3.92:96.08 wt.%), 350 °C for 10 min; post ion-exchange annealing — air atmosphere (600 °C for 45 h) Ion-exchange technique [330] [329] [328] [327] [326] Ref. (continued) Javier FONSECA. NPs embedded into glass matrices: GNCs 45 Borosilicate glass (72.0 wt.% SiO2, 14.0 wt.% Na2O, 0.6 wt.% K2O, 7.1 wt.% CaO, 4.0 wt.% MgO, 1.9 wt.% Al2O3, 0.1 wt.% Fe2O3, and 0.3 wt.% SO3) Borosilicate glass (72.0 wt.% SiO2, 14.0 wt.% Na2O, 0.6 wt.% K2O, 7.1 wt.% CaO, 4.0 wt.% MgO, 1.9 wt.% Al2O3, 0.1 wt.% Fe2O3, and 0.3 wt.% SO3) Borosilicate glass (72.0 wt.% SiO2, 14.0 wt.% Na2O, 0.6 wt.% K2O, 7.1 wt.% CaO, 4.0 wt.% MgO, 1.9 wt.% Al2O3, 0.1 wt.% Fe2O3, and 0.3 wt.% SO3) Borosilicate glass (72.0 wt.% SiO2, 14.0 wt.% Na2O, 0.6 wt.% K2O, 7.1 wt.% CaO, 4.0 wt.% MgO, 1.9 wt.% Al2O3, 0.1 wt.% Fe2O3, and 0.3 wt.% SO3) Cu NPs@silicate glass Cu NPs@silicate glass Cu NPs@silicate glass Cu NPs@silicate glass Ag NPs Silicate glass (69.59 wt.% SiO2, 15.17 wt.% Na2O, 5.08 wt.% MgO, 6.52 wt.% CaO, 1.73 wt.% Al2O3, 1.14 wt.% K2O, and 0.78 wt.% SO3) Silicate glass (69.59 wt.% SiO2, 15.17 wt.% Na2O, 5.08 wt.% MgO, 6.52 wt.% CaO, 1.73 wt.% Al2O3, 1.14 wt.% K2O, and 0.78 wt.% SO3) Ag NPs@silicate glass Ag NPs@silicate glass Ag NPs Ag NPs Soda-lime silicate glass (69.36 wt.% SiO2, 15.63 wt.% Na2O, 3.04 wt.% Al2O3, 6.05 wt.% CaO, 3.41 wt.% MgO, 1.72 wt.% K2O, 0.53 wt.% SO3, and 0.27 wt.% TiO2) Ag NPs@soda-lime silicate glass Ag NPs Soda-lime silicate glass (69.36 wt.% SiO2, 15.63 wt.% Na2O, 3.04 wt.% Al2O3, 6.05 wt.% CaO, 3.41 wt.% MgO, 1.72 wt.% K2O, 0.53 wt.% SO3, and 0.27 wt.% TiO2) Cu NPs Cu NPs Cu NPs Cu NPs Cu NPs Ag NPs Ag NPs@soda-lime silicate glass Cu NPs@silicate glass Ag NPs@silicate glass Ag NPs Ag NPs Silicate glass (5.73 wt.% Na2O, 7.52 wt.% ZnO, 9.42 wt.% Al2O3, 5.55 wt.% SiO2, 36.42 wt.% Yb2O3, and 35.36 wt.% Er2O3) Silicate glass (5.75 wt.% Na2O, 7.55 wt.% ZnO, 9.46 wt.% Al2O3, 5.58 wt.% SiO2, 36.58 wt.% Yb2O3, and 35.07 wt.% Ho2O3) Silicate glass (5.71 wt.% Na2O, 7.50 wt.% ZnO, 9.39 wt.% Al2O3, 5.54 wt.% SiO2, 36.31 wt.% Yb2O3, and 35.55 wt.% Tm2O3) Borosilicate glass (72.0 wt.% SiO2, 14.0 wt.% Na2O, 0.6 wt.% K2O, 7.1 wt.% CaO, 4.0 wt.% MgO, 1.9 wt.% Al2O3, 0.1 wt.% Fe2O3, and 0.3 wt.% SO3) Ag NPs@silicate glass Ag NPs@silicate glass Reinforcement Host matrix Composite (2.0) (1.6) 4‒6 2‒4 (10) (9.5) (8) (7) (7) 10‒40 10‒60 3‒10 Size of NPs (minmax (mean))/nm Thermal annealing in air: ion-exchange — CuSO4·5H2O:Na2SO4 melt (67.36:32.64 wt.%), 590 °C for 2 min; post ion-exchange annealing — air atmosphere (450 °C for 1 h) Thermal annealing in air: ion-exchange — CuSO4·5H2O:Na2SO4 melt (67.36:32.64 wt.%), 590 °C for 2 min; post ion-exchange annealing — air atmosphere (500 °C for 1 h) Thermal annealing in air: ion-exchange — CuSO4·5H2O:Na2SO4 melt (67.36:32.64 wt.%), 590 °C for 2 min; post ion-exchange annealing — air atmosphere (550 °C for 1 h) Thermal annealing in air: ion-exchange — CuSO4·5H2O:Na2SO4 melt (67.36:32.64 wt.%), 590 °C for 2 min; post ion-exchange annealing — air atmosphere (600 °C for 1 h) Thermal annealing in air: ion-exchange — CuSO4·5H2O:Na2SO4 melt (67.36:32.64 wt.%), 590 °C for 2 min; post ion-exchange annealing — air atmosphere (650 °C for 1 h) Thermal annealing in air: ion-exchange — AgNO3:NaNO3 melt (0.2:99.8 wt.%), 320 °C for 30 min; post ion-exchange annealing — H2 atmosphere (180 °C for 12 h) Thermal annealing in air: ion-exchange — AgNO3:NaNO3 melt (0.2:99.8 wt.%), 320 °C for 30 min; post ion-exchange annealing — H2 atmosphere (250 °C for 5 h) Thermal annealing in air: ion-exchange — AgNO3:NaNO3 melt (1.98:98.02 wt.%), 320 °C for 20 min; post ion-exchange annealing — air atmosphere (450 °C for 1 h) Thermal annealing in air: ion-exchange — AgNO3:NaNO3 melt (1.98:98.02 wt.%), 320 °C for 20 min; post ion-exchange annealing — air atmosphere (500 °C for 1 h) Thermal annealing in air: ion-exchange — AgNO3:NaNO3:KNO3 melt (14:45:41 wt.%), 280 °C for 50 min Ion-exchange technique ‒ NLO devices NLO devices Photoluminescence Application [333] [332] [331] [79] Ref. (continued) 46 Front. Mater. Sci. 2022, 16(3): 220607 Ag NPs Ag NPs Ag NPs Silicate glass (69.59 wt.% SiO2, 15.17 wt.% Na2O, 5.08 wt.% MgO, 6.52 wt.% CaO, 1.73 wt.% Al2O3, 1.14 wt.% K2O, and 0.78 wt.% SO3) Silicate glass (69.59 wt.% SiO2, 15.17 wt.% Na2O, 5.08 wt.% MgO, 6.52 wt.% CaO, 1.73 wt.% Al2O3, 1.14 wt.% K2O, and 0.78 wt.% SO3) Silicate glass (69.59 wt.% SiO2, 15.17 wt.% Na2O, 5.08 wt.% MgO, 6.52 wt.% CaO, 1.73 wt.% Al2O3, 1.14 wt.% K2O, and 0.78 wt.% SO3) Silicate glass (69.59 wt.% SiO2, 15.17 wt.% Na2O, 5.08 wt.% MgO, 6.52 wt.% CaO, 1.73 wt.% Al2O3, 1.14 wt.% K2O, and 0.78 wt.% SO3) Silicate glass (69.59 wt.% SiO2, 15.17 wt.% Na2O, 5.08 wt.% MgO, 6.52 wt.% CaO, 1.73 wt.% Al2O3, 1.14 wt.% K2O, and 0.78 wt.% SO3) Silicate glass (69.59 wt.% SiO2, 15.17 wt.% Na2O, 5.08 wt.% MgO, 6.52 wt.% CaO, 1.73 wt.% Al2O3, 1.14 wt.% K2O, and 0.78 wt.% SO3) Silicate glass (69.59 wt.% SiO2, 15.17 wt.% Na2O, 5.08 wt.% MgO, 6.52 wt.% CaO, 1.73 wt.% Al2O3, 1.14 wt.% K2O, and 0.78 wt.% SO3) Silicate glass (69.59 wt.% SiO2, 15.17 wt.% Na2O, 5.08 wt.% MgO, 6.52 wt.% CaO, 1.73 wt.% Al2O3, 1.14 wt.% K2O, and 0.78 wt.% SO3) Ag NPs@silicate glass Ag NPs@silicate glass Ag NPs@silicate glass Ag NPs@silicate glass Ag NPs@silicate glass Ag NPs@silicate glass Ag NPs@silicate glass Ag NPs@silicate glass Ag NPs Ag NPs Ag NPs Ag NPs Ag NPs Reinforcement Host matrix Composite (1.9) (2.2) (2.6) (2.4) (2.6) (2.7) (2.7) (2.6) Size of NPs (minmax (mean))/nm Thermal annealing in air: ion-exchange — AgNO3:NaNO3 melt (1.98:98.02 wt.%), 320 °C for 20 min; post ion-exchange annealing — air atmosphere (550 °C for 1 h) Irradiation: ion-exchange — AgNO3:NaNO3 melt (1.98:98.02 wt.%), 320 °C for 20 min; post ion-exchange annealing — air atmosphere (550 °C for 1 h); post ion-exchange irradiation (λ = 266 nm, number of pulses = 50, laser fluence = 18 J·cm−2) Irradiation: ion-exchange — AgNO3:NaNO3 melt (1.98:98.02 wt.%), 320 °C for 20 min; post ion-exchange annealing — air atmosphere (550 °C for 1 h); post ion-exchange irradiation (λ = 266 nm, number of pulses = 200, laser fluence = 72 J·cm−2) Irradiation: ion-exchange — AgNO3:NaNO3 melt (1.98:98.02 wt.%), 320 °C for 20 min; post ion-exchange annealing — air atmosphere (550 °C for 1 h); post ion-exchange irradiation (λ = 355 nm, number of pulses = 40, laser fluence = 14.4 J·cm−2) Irradiation: ion-exchange — AgNO3:NaNO3 melt (1.98:98.02 wt.%), 320 °C for 20 min; post ion-exchange annealing — air atmosphere (550 °C for 1 h); post ion-exchange irradiation (λ = 355 nm, number of pulses = 200, laser fluence = 72 J·cm−2) Irradiation: ion-exchange — AgNO3:NaNO3 melt (1.98:98.02 wt.%), 320 °C for 20 min; post ion-exchange annealing — air atmosphere (550 °C for 1 h); post ion-exchange irradiation (λ = 532 nm, number of pulses = 10, laser fluence = 3.6 J·cm−2) Irradiation: ion-exchange — AgNO3:NaNO3 melt (1.98:98.02 wt.%), 320 °C for 20 min; post ion-exchange annealing — air atmosphere (550 °C for 1 h); post ion-exchange irradiation (λ = 532 nm, number of pulses = 50, laser fluence = 18 J·cm−2) Irradiation: ion-exchange — AgNO3:NaNO3 melt (1.98:98.02 wt.%), 320 °C for 20 min; post ion-exchange annealing — air atmosphere (550 °C for 1 h); post ion-exchange irradiation (λ = 532 nm, number of pulses = 200, laser fluence = 72 J·cm−2) Ion-exchange technique ‒ ‒ Application [333] [333] Ref. (continued) Javier FONSECA. NPs embedded into glass matrices: GNCs 47 Ag NPs Soda-lime silicate glass (74.2 wt.% SiO2, 14.3 wt.% Na2O, 1.9 wt.% Al2O3, 8.1 wt.% CaO, and 1.5 wt.% MgO) Silicate glass (71.92 wt.% SiO2, 13.31 wt.% Na2O, 4.15 wt.% MgO, 8.70 wt.% CaO, and 1.92 wt.% Al2O3) Ag NPs@soda-lime silicate glass Ag NPs@soda-lime silicate glass Soda-lime silicate glass Ag NPs Soda-lime silicate glass (74.2 wt.% SiO2, 14.3 wt.% Na2O, 1.9 wt.% Al2O3, 8.1 wt.% CaO, and 1.5 wt.% MgO) Ag NPs@soda-lime silicate glass Ag NPs@silicate glass Ag NPs Soda-lime silicate glass (74.2 wt.% SiO2, 14.3 wt.% Na2O, 1.9 wt.% Al2O3, 8.1 wt.% CaO, and 1.5 wt.% MgO) Ag NPs@soda-lime silicate glass Ag NPs Ag NPs Reinforcement Host matrix Composite (4.2) (4.93 ± 1.88) (2) (1) 3‒8 (7) Size of NPs (minmax (mean))/nm Thermal annealing in air: ion-exchange — AgNO3:NaNO3 melt (3.92:96.08 wt.%), 320 °C for 10 min; post ion-exchange annealing—air atmosphere (600 °C for 45 h) Thermal annealing in air: ion-exchange — AgNO3:NaNO3 melt (3.92:96.08 wt.%), 320 °C for 10 min; post ion-exchange annealing — UV-laser irradiation, λ = 193 nm, pulse energy density (30 mJ∙cm−2), repetition frequency (10 Hz), pulse duration (20 ns), irradiation time (5 s) Thermal annealing in air: ion-exchange — AgNO3:NaNO3 melt (3.92:96.08 wt.%), 320 °C for 10 min; post ion-exchange annealing — UV-laser irradiation, λ =193 nm, pulse energy density (30 mJ∙cm-2), repetition frequency (10 Hz), pulse duration (20 ns), irradiation time (30 min) Irradiation: ion-exchange — AgNO3:NaNO3 melt (20:80 wt.%), 350 °C for 30 s; post ion-exchange irradiation — Ar+ ions of 200 keV, ion fluence (5 × 1015 ion·cm−2), flux (2 µA·cm−2) Irradiation: ion-exchange — AgNO3:NaNO3 melt (2:98 wt.%), 400 °C for 2h; post ion-exchange irradiation — 100 kV electrons, flux (6.4 A·cm−2) Ion-exchange technique ‒ Optoelectronics ‒ ‒ Application [336] [335] [334] [334] Ref. (continued) 48 Front. Mater. Sci. 2022, 16(3): 220607 Javier FONSECA. NPs embedded into glass matrices: GNCs the glass and the electron donor affinity of the O2− ions [344]. Ag NPs have also been incorporated in a silicate glass (SiO2‒Al2O3‒Na2O‒MgO‒ZnO) through the ionexchange process [325]. The Ag+ ions replaced the smaller the Na+ ions in the glass host during the Ag+‒Na+ ion-exchange. Before the formation of NPs, a portion of Ag+ ions was reduced to Ag0 atoms as follows: 2(≡SiO− ) + Ag+ ↔ ≡Si − O − O − Si≡ +Ag0 (34) The subsequent formation of Ag NPs was suggested to proceed via Ostwald ripening [345‒347]. The influence of ion-exchange parameters on the size of the Ag NPs was investigated (Fig. 25): (1) The silicate glass was subjected to Ag+‒Na+ ionexchange using a molten mixture of AgNO3 and NaNO3 with different molar ratios at 350 °C for 1 h. The size and quantity of Ag NPs were found to increase as the AgNO3:NaNO3 ratio increased. When the AgNO3:NaNO3 ratio was 1:99, 7:93, 20:80, and 50:50 mol.%, the mean size of the Ag NPs was 1.5, 2.0, 2.2, and 2.3 nm, respectively. (2) The silicate glass was subjected to Ag+‒Na+ ionexchange at 380 °C for 15, 90, and 240 min. The AgNO3:NaNO3 ratio was fixed at 7:93 mol.%. The average size of the Ag NPs was approximately 1.6, 2.0, and 2.5 nm when the ion-exchange time was 15, 90, and 240 min, respectively. Therefore, the size of the Ag NPs increased as the ion-exchange time increased. (3) The silicate glass was subjected to Ag+‒Na+ ionexchange at different temperatures (300, 350, and 400 °C) for 1 h. The AgNO3:NaNO3 ratio was also fixed at 7:93 mol.%. The mean size of the Ag NPs was 1.8, 2.3, and 2.4 nm when the ion-exchange temperature was 300, 49 350, and 400 °C, respectively. Therefore, increasing the ion-exchange temperature resulted in the formation of larger Ag NPs. (4) The glass was doped with Eu3+ ions. The Eu3+doped glass was subjected to Ag+‒Na+ ion-exchange using a variable AgNO3:NaNO3 ratio at 380 °C for 1 h. The average size of the Ag NPs was 3.0, 3.3, 3.4, and 3.6 nm when the AgNO3:NaNO3 ratio was 1:99, 7:93, 20:80, and 50:50 mol.%, respectively. Therefore, under the same preparation conditions, the Ag NPs within the Eu3+-doped glass were larger than those embedded in the glass without Eu3+ ions. The formation of larger Ag NPs was attributed to the self-reduction of Eu3+ ions to Eu2+ ions which proceeded as follows [195,348]: Eu3+ + e− → Eu2+ (35) Eu2+ + Ag+ → Eu3+ + Ag0 (36) zAg0 → Agz (37) It is worth mentioning that the Ag NPs enhanced the luminescence of the embedded Eu3+ ions. The energy efficient transfer from Ag NPs to Eu3+ ions was suggested to be the reason for the luminescence enhancement of Eu3+ [325]. Chen et al. and Karvonen et al. have embedded Ag NPs into a commercial glass (Corning 0211) by an unconventional two-step ion-exchange method [326,349]. Corning 0211 glass was placed in a molten mixture of AgNO3:NaNO3:KNO3 (5:47.5:47.5 mol.%) at 300 °C for 6 h. In doing so, ion-exchange occurred. Then, an Al film was evaporated on the ion-exchanged glass using an electron-beam evaporator. The Al film provided electrons Fig. 25 Schematic illustration of the formation and growth of Ag NPs. Reproduced with permission from Ref. [325] (Copyright 2018 Elsevier). 50 Front. Mater. Sci. 2022, 16(3): 220607 to reduce Ag+ ions to Ag0 atoms. The GNC was immersed in a KNO3 salt at 400 °C for 2 h to promote the formation of Ag NPs. The KNO3 salt favored the flow of Ag+ ions during the electrolytic deposition. The average size of the embedded Ag NPs was found to be about 10 nm. Moreover, the size distribution of Ag NPs was small. At the wavelength of 800 nm, this GNC exhibited n2 and β of 1.4 × 10−10 cm2·W−1 and −1.9 × 10−5 cm·W−1, respectively [326]. 3.4.3 Thermal annealing in air After ion-exchange process, the most common and simple strategy for precipitating metal NPs within a glass matrix is thermal annealing performed in air at a temperature close to the glass transition temperature. Thermal annealing produces free electrons from NBOs in the glass matrix. These electrons are captured by metal cations to form metal atoms. The formation and growth of metal NPs generally increases as the annealing time and temperature increase (as long as the temperature remains below the glass transition temperature) [350]. Mathpal research’s group has explored the incorporation Ag NPs into a soda-lime silicate glass (SiO2‒Na2O‒CaO‒MgO‒Al2O3‒K2O‒SO3‒Fe2O3) by Ag+‒Na+ ion-exchange followed of thermal annealing in an air atmosphere [327‒328]. The glass was immersed in a molten salt bath of AgNO3:NaNO3 (7:93 mol.%) at 370 °C for 2 min. Under these conditions, the Ag+ ions diffused into the glass matrix and replaced the Na+ ions in the glass. Then, the ion-exchanged composite was annealed in an oxidizing atmosphere at 450, 500, and 550 °C for 1 h. The ion-exchange generated stress in the system due to the difference in size of the Ag+ and Na+ ions. This stress was relaxed by the diffusion of Ag+ ions near the surface of the soda-lime glass during the thermal annealing. Thus, the total energy of the system was minimized. Annealing also resulted in the reduction of Ag+ ions to Ag0 atoms and the subsequent formation of spherical Ag NPs. Consequently, most of the Ag NPs were located near the surface of the soda-lime glass. The average size of the embedded Ag NPs was (2.8 ± 1), (3.3 ± 1.2), and (4.0 ± 1.5) nm for the GNCs annealed at 450, 500, and 550 °C, respectively [327]. The soda-lime silicate glass was also immersed in a mixture of AgNO3:NaNO3 (20:80 mol.%) at 370 °C for 2 min [328]. This ion-exchanged composite was annealed at 500, 550, and 600 °C for 1 h. The size of the Ag NPs was found to be 2.9, 4.4, and 7.2 nm when the annealing temperature was 500, 550, and 600 °C, respectively. The Ag NPs also accumulated near the surface of the soda-lime glass during thermal annealing. The largest accumulation of Ag NPs near the glass surface was found in the GNC annealed at 550 °C. The near surface accumulation in the GNC annealed at 600 °C was between that of the GNC annealed at 500 °C and that of the GNC annealed at 550 °C. Therefore, when the annealing temperature was 600 °C, the Ag0 atoms diffused deep into the glass matrix before aggregation to form Ag NPs [328]. Regardless of the AgNO3:NaNO3 molar ratio of the molten bath, the size of the NPs always increased as the annealing temperature increased [327‒328]. The PL intensity of these GNCs was reported to decrease as the temperature and time of the post ion-exchange annealing treatment increased. Therefore, the growth of Ag NPs resulted in the quenching of PL intensity [327‒328]. The formation mechanism of Ag NPs within the sodalime silicate glass (SiO2‒Al2O3‒Fe2O3‒TiO2‒SO3‒CaO‒ MgO‒Na2O‒K2O) was explored by Simo et al. [329]. Ag NPs were embedded within this glass by combining ionexchange and subsequent annealing in air (Fig. 26). The ion-exchange was carried out by immersing the glass in a mixture of AgNO3:NaNO3 (33.33:66.66 mol.%) at 360 °C for 1 h. Na+ ions from the surface and the subsurface region of the glass were replaced by Ag+ ions. Ag clusters with a size less than 1 nm in diameter (approximately 0.6 nm) were generated during ion-exchange. The formation of NPs was initiated by heat treatment. This post ion-exchange treatment was carried out at different temperatures and times: (1) The ion-exchanged glass was annealed in air at different temperatures (between 100 and 600 °C) for 30 min. Ag clusters (d < 1 nm) persisted at annealing temperatures below 410 °C. Ag NPs were formed above 410 °C. It was suggested that Ag+ ions were reduced by electrons extracted directly from NBOs as follows [351‒352]: 2(≡Si−O− Ag+ ) ↔ ≡Si−O−Si≡ +O− + (Ag2 )+ O− + (Ag2 )+ → 2Ag0 + 0.5O2 (38) (39) The size of the Ag NPs was found to increase slightly as the annealing temperature increased. The average size of the Ag NPs formed was (1.8 ± 1.1), (1.9 ± 1.2), and (2.2 ± 0.6) nm at 450, 500, and 600 °C, respectively. Javier FONSECA. NPs embedded into glass matrices: GNCs 51 Fig. 26 Schematic illustration of the preparation of Ag NPs within a soda-lime silicate glass by ion-exchange and subsequent annealing. Reproduced with permission from Ref. [329] (Copyright 2012 American Chemical Society). (2) Annealing was performed at 400 °C for times ranging from 1 min to 13 h. Ag NPs were not formed under these conditions. Only the number of Ag cluster increased as the annealing time increased. The composite annealed for 13 h at 400 °C contained exclusively Ag clusters. Moreover, the Ag NPs were also not formed when this annealed composite was subsequently treated at 500 °C. In other words, the Ag clusters were stable at high temperatures. Therefore, the Ostwald ripening mechanism did not happen. (3) The ion-exchanged glass was annealed at 550 °C for times ranging from 1 min to 16 h. The Ag clusters became Ag NP. Although the size of the Ag NPs increased, their number remained constant. Therefore, in contrast to the widely accepted Ostwald ripening processes of particle growth in this GNCs, the particle growth was produced by the addition of Ag monomers due to the decomposition of ) ( ≡ Si − O− Ag+ . In summary, the size of the Ag NPs depended on the concentration of the Ag cluster, which served as nuclei for the formation of Ag NPs, and on the duration of annealing. However, the size was limited by the number of reducing agents (NBOs). The ion-insertion process and the growth of NPs were separated in this process which is essential to control the size of the NPs [329]. In a very similar study, Sheng et al. have formed Ag NPs within a soda-lime silicate glass (SiO2‒Na2O‒Al2O3‒ CaO‒MgO) by Ag+‒Na+ ion-exchange and thermal annealing in air [330]. Firstly, the glass was dipped in a molten mixture of AgNO3:NaNO3 (2:98) at different temperatures (320 or 350 °C) for 10 min. This ionexchange did not generate Ag NPs but Ag clusters (< 1 nm). The ion-exchanged glasses were then annealed in air at 570 or 600 °C for up to 100 h. The formation of Ag NPs occurred after heating at 570 or 600 °C for more than 25 h. The mean size of the Ag NPs was approximately 2 nm after annealing at 600 °C for 45 h. The Ag NPs grew more as the temperature and time of annealing increased. After ion-exchange, Ag+ ions were mainly attached to NBO atoms. During annealing, the Ag−O bonds were broken. The resulting Ag+ ions diffused to the glass surface to minimize the surface tensile stress introduced by the size difference between Ag+ and Na+ during the ion-exchange process. Finally, Ag+ ions were reduced to Ag0 atoms by capturing electrons from the glass structure and defects during annealing [330]. Ag NPs have also been incorporated in a photo-thermorefractive (PTR) glass (SiO2‒Na2O‒ZnO‒Al2O3‒F‒Sb2O3) through ion-exchange and subsequent heat treatment in air [88,353]. It should be mentioned that PTR glasses are multicomponent glasses whose n2 can vary with UV irradiation and heat treatment due to the growth of NaF NCs in the UV irradiated areas [354]. These glasses are widely used in photonics [355]. The glass was immersed in a melt bath of AgNO3:NaNO3 (5:95 mol.%) at 320 °C for different durations (from 5 min to 21 h). The ionexchanged glass was treated at different temperatures and durations. The heat treatment of the glass at 450 °С induced: (i) diffusion of Ag+ ions towards the glass surface, (ii) reduction of Ag+ ions to Ag0 atoms, and (iii) aggregation of isolated Ag0 atoms in NPs. The formation of Ag clusters and NPs was reported to be adjusted by controlling the temperature and time of the post ionexchange treatment. The Sb3+ ions, which were contained in the PTR glass, were found to act as electron donors for the Ag+ ions. It was also observed that an increase in Sb3+ ions in the PTR glass increased the rate of reduction of Ag+ ions to Ag0 atoms, which is consistent with the law of mass action: 52 Front. Mater. Sci. 2022, 16(3): 220607 2Ag+ + Sb3+ ↔ 2Ag0 + Sb5+ (40) Therefore, the formation of Ag NPs was highly dependent on the concentration of Sb3+ ions in the initial PTR glass. In fact, Ag+ ions remained in ionic form within a PTR glass without Sb3+ ions regardless of the post ionexchange treatment temperature [88,353]. Varak et al. have incorporated Ag NPs into RE-doped silicate glasses by ion-exchange and subsequent heat treatment in air [79]. First, three silicate glasses (Na2O‒ZnO‒Al2O3‒SiO2‒Yb2O3‒RE2O3, where RE = Er, Ho, Tm) were prepared by melt-quenching technique. Then, Ag+ ions were incorporated into the glasses by ionexchange. The AgNO3:NaNO3:KNO3 melt (14:45:41 wt.%) was used for ion-exchange at 280 °C for 50 min. Finally, the ion-exchanged glasses were annealed in air at 590 °C for 1 h. Before annealing, the Ag+ ions remained in a surface layer. However, when annealing was applied, the Ag+ ions were reduced and formed Ag dimers and NPs that migrated deeper into the glasses. These Ag species were found in a layer 8 to 15 µm deep. The largest amount of Ag NPs was present in the Ho-doped glass. The size of these Ag NPs ranged between 10 and 60 nm. The size of Ag NPs embedded within Er and Tm doped glasses was found to vary from 3 to 10 nm and 10 to 40 nm, respectively. It is worth mentioning that Ag species enhanced the PL of the RE ions embedded within the glass matrices by various mechanisms. Upon excitation in the 200 to 450 nm range, Ag dimers increased the PL of RE-doped glasses by the energytransfer mechanism. After excitation at 975 nm, the PL of the RE-doped glasses increased due to the SPR effect of Ag NPs and due to structural changes of the silicate matrices, which were generated by Na+ ↔ Ag+ substitution during ion-exchange [79]. Ag NPs have been embedded in a Eu3+-doped sodium‒aluminosilicate glass (SiO2‒Al2O3‒Na2O‒MgO‒ ZnO‒Eu2O3) by Ag+‒Na+ ion-exchange and subsequent heat treatment [195]. The glass was immersed in a molten mixture of AgNO3:NaNO3 (5:95 mol.%) at 360 °C for 2 h. Then, the ion-exchanged glass was annealed at 450 °C for 10 h. Eu3+ ions spontaneously reduced to Eu2+ in the glass matrix [356‒357]. As the standard reduction potential (E0) of Ag+/Ag0 is higher (0.7996 eV) than that of Eu3+/Eu2+ (−0.35 eV), it was assumed that reaction (Eq. (41)) might have occurred in the Ag+ ion-exchange process: Eu2+ + Ag+ → Eu3+ + Ag0 (41) Subsequently, the following reactions might have taken place: Ag0 + Ag+ → Ag+2 (42) zAg0 → Agz (43) This can explain that the concentration of Ag NPs and ionic Ag species was found to increase as the concentration of Eu3+ ions increased [195]. Beyond Ag NPs, other metal NPs have also been incorporated into glass matrices by ion-exchange and post-treatment. For instance, Cu NPs have been precipitated in a soda-lime silicate glass (SiO2‒Na2O‒ K2O‒CaO‒MgO‒Al2O3‒Fe2O3‒SO3) [331]. The glass was immersed in a mixture of CuSO4·5H2O:Na2SO4 (54:46 mol.%) at 590 °C for 2 min to perform Cu+‒Na+ ion-exchange. The embedded Cu+ ions were subsequently reduced to Cu0 atoms by heat treatment in air at 450, 500, 550, 600, and 650 °C for 1 h. These Cu0 atoms aggregated near the surface of the glass to form evenly distributed Cu NPs. The average size of the Cu NPs within the GNC treated at 450, 500, 550, 600, and 650 °C was (7 ± 1), (7 ± 2), (8 ± 3), (9.5 ± 1), and (10 ± 2) nm, respectively (Fig. 27). In other words, the size of the Cu NPs increased as the temperature for heat treatment increased. Therefore, the nucleation and growth of Cu NPs were controlled by post ion-exchange heat treatment. These GNC, regardless of the heat treatment temperature, exhibited good NLO behavior with possible applications in the field of laser technology and all-optical switching devices [331]. Different metal NPs have been embedded together in a glass matrix. Cu NPs and Ag NPs have been incorporated in a same soda-lime glass (SiO2‒Na2O‒CaO‒MgO‒ Al2O3‒SO3‒Fe2O3) by sequential Cu+‒Na+ and Ag+‒Na+ ion-exchange and subsequent heat treatment in air [358]. The glass was dipped in a molten salt bath of CuSO4·5H2O:Na2SO4 for about 1 min and then in a molten salt bath of AgNO3:NaNO3 also for also about 1 min. Annealing was performed in air at different temperatures ranging from 250 to 650 °C for 1 h. Cu NPs and Ag NPs were formed within the glass matrix. At high annealing temperatures, Cu NPs were found to dissolve during Ag clusterization. When the glass matrix was first Ag+‒Na+ ion-exchanged, then Cu+‒Na+ ion-exchanged, and finally annealed, only the Ag NPs precipitated within the glass matrix [358]. Javier FONSECA. NPs embedded into glass matrices: GNCs 53 Fig. 27 TEM images and corresponding Ag NP size distributions of the GNC treated (after ion-exchange) at (a)(b) 450 °C, (c)(d) 500 °C, (e)(f) 550 °C, (g)(h) 600 °C, and (i)(j) 650 °C. Reproduced with permission from Ref. [331] (Copyright 2019 Elsevier). 3.4.4 Thermal annealing in H2 atmosphere Another post ion-exchange treatment of a glass is thermal annealing in a controlled H2 atmosphere. This post ionexchange treatment not only favors the formation of NPs within the glass but also the formation of metal island films (MIFs) on the glass surface. MIFs are generally produced on the glass surface by out-diffusion of metal atoms due to the hydrogen penetration into the glass [332,359]. This technique requires safety precautions. In an early example, Ag NPs have been embedded into a soda-lime silicate glass (SiO2‒Na2O‒Al2O3‒CaO‒ MgO‒K2O‒SO3‒TiO2) by Ag+‒Na+ ion-exchange and subsequent annealing in a H2 atmosphere [332]. The glass was immersed in a molten mixture of AgNO3:NaNO3 (0.1:99.9 mol.%) at 320 °C for 30 min. About 20% of the Na+ ions were replaced by Ag+ ions. The ion-exchanged glasses were treated in H2 at different temperatures (120, 140, 160, 180, 200, and 250 °C) for 5 h. The ionexchanged glasses were also annealed in H2 at 180 °C for different times (2, 4, 8, and 12 h). The annealing in H2 atmosphere was found to cause the diffusion of Ag+ ions towards the glass surface and also the near-surface precipitating of Ag NPs. The permeation of H2 in the glass was accompanied by the Na+‒H+ ion-exchange. It is worth mentioning that some Na+ ions remained in the glass matrix after the Ag+‒Na+ ion-exchange. The migration of Ag+ ions towards the glass surface was attributed to a direct interaction with H2, but also to a charge balance during Na+‒H+ ion-exchange. The size of the Ag NPs within the glasses annealed in H2 at 180 °C for 12 h and at 250 °C for 5 h ranged between 2 and 4 nm and between 4 and 6 nm, respectively. The size of the Ag NPs was reported to increase as the annealing time and temperature increased. Furthermore, to obtain a direct comparison of the effect of the annealing atmosphere, an ion-exchanged glass was treated in air at 200 °C for 5 h. Ag NPs were not found within this composite. Therefore, H2 annealing was carried out at temperatures low enough to avoid NP formation from thermal effects alone [332]. 3.4.5 Laser irradiation techniques Laser irradiation can also be used as post ion-exchange 54 Front. Mater. Sci. 2022, 16(3): 220607 treatment of glass matrices for the precipitation of metal NPs within such glasses. Continuous [360‒363] and/or pulsed lasers [364‒366] working at different wavelengths of the spectrum can reduce metal ions to metal NPs. The formation of NPs depends on the laser power at the surface of the ion-exchanged glass matrix. When the laser power ensures that the temperature of the irradiated area remains close to the glass transition temperature throughout the process, the interaction of the laser beam with the sample is thermal in nature. In other words, the reduction of metal cations follows a similar trend to that of thermal annealing. However, this post treatment is faster in terms of NP formation [350]. At higher laser powers, the glass surface temperature is higher than the glass transition temperature. Consequently, the irradiated region of the ion-exchanged glass melts. In this liquid state, NPs generally form in a deep region of the irradiated area and migrate towards the edges of the laser beam where they aggregate to form larger NPs (Fig. 28) [360‒362]. Rahman et al. have studied the precipitation of Ag NPs in Ag+‒Na+ ion-exchanged soda-lime silicate glass (SiO2‒Na2O‒MgO‒CaO‒Al2O3‒K2O‒SO3) by different post ion-exchange treatments [333]. The glass was immersed in a molten salt bath of AgNO3:NaNO3 (1:99 mol.%) at 320 °C for 20 min to perform Ag+‒Na+ ion-exchange. Subsequently, the ion-exchanged sample was annealed at different temperatures (450, 500, and 550 °C) for 1 h and irradiated at different wavelengths (266, 355 and 532 nm) and different number of consecutive laser pulses. Ag+ ions were reduced to Ag NPs during annealing in air. The Ag NPs were progressively larger as the annealing temperature increased. A strong interaction between the light and the Fig. 28 Formation and growth of metal NPs by high power laser irradiation. Ag NPs appeared when the 550 °C annealed composite was irradiated at 532 nm. Consequently, ionized multimeric structures, such as (Ag3)2+, were formed from the Ag NPs. Thus, the size of Ag NPs decreased. Similarly, the size of the Ag NPs within the 550 °C annealed composite was reduced by irradiation with many laser shots (200 pulses) at 355 nm. Ag+ ions were released from the Ag NPs into the glass matrix. This resulted in the depolymerization of the silica network. Under 266 nm laser irradiation on the 550 °C annealed sample, the temperature gradient between the Ag NPs and the surrounding glass matrix remained low. There was no size reduction of Ag NPs. Laser irradiation at this wavelength was found to promote the formation of Ag NPs with very narrow size distribution. In brief, the size and size distribution of Ag NPs embedded in annealed glass matrices can be adjusted by selecting laser irradiation wavelength and power density (Fig. 29). Furthermore, Rahman et al. also studied the formation and precipitation of Ag NPs within soda-lime silicate glass solely by laser irradiation. This is considered a multiphoton process [367]. Valence band electrons are excited to the conduction band by absorbing photons. Free electrons in the conduction band react with Ag+ ions, reducing them to Ag0 atoms. However, according to the results of Rahman et al., irradiation with a 532 nm (2.32 eV) laser on Ag+‒Na+ ion-exchanged soda-lime silicate glass did not generate free charge carriers in SiO2 (bandgap ≈ 8.9 eV). The reduction of Ag+ and the subsequent precipitation of Ag NPs under laser irradiation with 532 nm on Ag+‒Na+ ion-exchanged soda-lime silicate glass were described as follows: ≡Si − O − Ag+ + hv → ≡Si − O − h+ + (e− + Ag+ ) (44) e− + Ag+ → Ag0 (45) zAg0 → Agz (46) Electrons were driven out from the 2p orbital of NBO in the glass and captured by nearby Ag+ ions to form Ag0 atoms, which then precipitated Agz. When Ag+‒Na+ ionexchanged soda-lime silicate glass was treated with 266 nm (4.66 eV) laser irradiation, the free electrons could have been driven out by both mechanisms: the multiphoton process and the 2p orbital of NBO. Therefore, the 266 nm laser irradiation is a more efficient irradiation than other wavelengths in reducing Ag+ ions Javier FONSECA. NPs embedded into glass matrices: GNCs 55 Fig. 29 Schematic illustration of the evolution of Ag species as a function of the different post ion-exchange treatments. Reproduced with permission from Ref. [333] (Copyright 2021 Elsevier). and promoting Ag NP precipitation [333]. Zhang et al. have also explored different post-treatments to precipitate Ag NPs in an ion-exchanged soda-lime silicate glass (SiO2‒Na2O‒Al2O3‒CaO‒MgO) [334]. The glass was immersed in a molten salt bath of AgNO3:NaNO3 (2:98 mol.%) at 320 °C for 10 min. The ion-exchanged glass was subject to (1) thermal annealing, or (2) UV-laser irradiation, or (3) X-ray irradiation to promote the formation of Ag NPs. (1) Thermal annealing was carried out in air at temperatures ranging between 500 and 600 °C for different durations (2‒45 h). Through annealing, Ag+ ions first diffused to the glass surface trying to minimize the surface tensile stress introduced during the ion-exchange process. Then, the Ag+ ions were reduced to Ag0 atoms by capturing the electron from the glass structure or impurities. Finally, the Ag0 atoms aggregated and became Ag NPs. When the ion-exchanged glass was annealed at 600 °C for 45 h, the size of the Ag NPs was found to vary between 3 and 8 nm with an average size of 7 nm. (2) The ion-exchanged glass was irradiated with an UVlaser (λ = 193 nm) with a pulse energy density of 30 mJ·cm−2. The UV-laser irradiation time ranged from 5 s to 30 min. The repetition frequency and pulse duration were 10 Hz and 20 ns, respectively. Electrons were driven out from the NBOs after laser irradiation. Laser Glass −−−−→ h+ + e− Ag+ ions captured those electrons to form Ag0 atoms. (47) 56 Front. Mater. Sci. 2022, 16(3): 220607 Ag+ + e− → Ag0 (48) The aggregation of Ag0 atoms was suggested to occur due to the high temperature of the glass. It is worth noting that UV-laser irradiation was responsible for the increase in temperature of the glass. Therefore, this irradiation promoted the formation of Ag NPs. After 5 s and after 30 min of irradiation, the average size of the Ag NPs was 1.0 and 2.0 nm, respectively. The size of the NPs increased as the UV-laser irradiation time increased. (3) The ion-exchanged glass was also treated with X-ray irradiation (Rh Kα, λ = 0.061 nm, 50 kV, 50 mA) at room temperature for 30 min. X-ray irradiation caused defects and induced the reduction of Ag0 atoms. However, it did not promote the formation of Ag NPs [334]. Ag NPs have also been precipitated within an Ag+‒Na+ ion-exchanged soda-lime silicate glass (SiO2‒Na2O‒ MgO‒CaO‒Al2O3) by irradiation with Ar at an incidence angle of 40° [335,368]. The soda-lime silicate glass was immersed in a molten mixture of AgNO3:NaNO3 (20:80 wt.%) at 350 °C for 30 s. In doing so, some Na+ ions within the glass were replaced by Ag+ ions contained in the molten mixture. To form and grow the Ag NPs, Ag+‒Na+ ion-exchanged soda-lime silicate glass was bombarded by 200 keV Ar+ ions (ion fluences = 5 × 1015 ions·cm−2; flux = 2 µA·cm−2) at an angle of incidence of 40° (Fig. 30). The 200 keV Ar+ ionized the Ag+‒Na+ ion-exchanged glass, thus providing electrons. Ag+ was reduced to Ag0 atoms by capturing those electrons. In addition, the energetic beam also created NBO defect centers [369‒370]. Ag0 atoms diffused to these NBO sites, where they accumulated and formed Ag NPs [328,370]. The Ag NPs were well distributed within the Ag+‒Na+ ion-exchanged glass matrix. The size of the Ag NPs ranged between 2 and 10 nm. The average size was found to be (4.93 ± 1.88) nm. Increasing Ar+ fluence from 5 × 1015 to 1 × 1016, and 5 × 1016 ions·cm−2 was found to convert more Ag+ ions to Ag0 atoms, which in turn diffused and agglomerated to form more Ag NPs. This increase in the concentration of Ag NPs decreased the intensity of PL and increased the electrical conductivity of the GNC [335]. Hofmeister et al. have embedded Ag NPs into the sodalime silicate glass by ion-exchange and subsequent electron beam irradiation [336]. Ion-exchange was carried out by immersing the glass in a molten mixture of AgNO3:NaNO3 (2:98 wt.%) at 400 °C for 2 h. Subsequently, the ion-exchanged glass was irradiated with Fig. 30 Schematic illustration of the preparation of Ag NPs within an Ag+‒Na+ ion-exchanged soda-lime silicate glass. Reproduced with permission from Ref. [335] (Copyright 2021 Elsevier). 100 kV electrons with a beam current density of 6.4 A·cm−2. During irradiation, the temperature of the glass was kept below 300 °C. The diffusivity of Ag+ ions achieved by irradiation was larger than that achieved by heat treatment. This, together with the bond breaking and the creation of defects within the glass network, favored the precipitation of high concentration of Ag NPs. The size distribution of the NPs was narrow. Ag NPs with a mean size of 4.2 nm were found to be homogeneously distributed within the glass matrix [336]. Therefore, the electron beam irradiation post-treatment may be superior to the heat post-treatment in the concentration of NPs produced, the homogeneous distribution of NPs within the glass matrix, the narrow size distribution of NPs, the speed of NPs formation and the reproducibility of the process. 3.5 Less conventional techniques Table 6 shows examples of GNCs prepared by staining process, spark plasma sintering, radio frequency sputtering, spray pyrolysis, and chemical vapor deposition [371‒376]. 3.5.1 Staining process The staining process is an especial ion-exchange method widely used in the fields of art and craft for coloring glasses. This method consists of enameling a glass surface with a mixture of a metal salt and a pigment. A subsequent annealing favors the ion-exchange process (metal cations — K+ or Na+). The metal cations are Silica glass Silica glass Silica glass Silica glass Silica glass Silica glass Silica glass Silica glass Silica glass GaAs NPs@silica glass GaAs NPs@silica glass GaAs NPs@silica glass Au NPs@silica glass Au NPs@silica glass Au NPs@silica glass Au NPs@silica glass Au NPs@silica glass Cu NPs@silica glass Au NPs@silica glass Cu NPs Au NPs Au NPs Au NPs Au NPs Au NPs GaAs NPs GaAs NPs GaAs NPs Au NPs Cu NPs Ag NPs Aluminosilicate glass (75.63 wt.% SiO2, 8.83 wt.% Na2O, 7.92 wt.% CaO, 4.22 wt.% MgO, 1.81 wt.% SnO2, 0.74 wt.% Al2O3, 0.43 wt.% Fe2O3, 0.27 wt.% SO3, and 0.14 wt.% K2O) Aluminosilicate glass (77.82 wt.% SiO2, 6.52 wt.% Na2O, 7.50 wt.% CaO, 4.24 wt.% MgO, 1.88 wt.% SnO2, 0.76 wt.% Al2O3, 0.03 wt.% Fe2O3, 0.25 wt.% SO3, and 1.00 wt.% K2O) Silica glass Ag NPs@ aluminosilicate glass Cu NPs@ aluminosilicate glass Reinforcement 14 47.08 17.5 8.09 3.64 2.43 7.7 4.4 2.7 3‒10 25‒40 2.5‒6 Size of NPs (min‒max (mean))/nm ‒ ‒ ‒ ‒ ‒ ‒ ‒ ‒ ‒ 0.05 ‒ ‒ Loading of NPs (wt.% in excess) Radio frequency sputtering: 13.56 MHz radio frequency sputtering, 4.7 mTorr Ar atmosphere at room temperature, radio frequency power (15 W), sputtering time (30 s) Radio frequency sputtering: 13.56 MHz radio frequency sputtering, 4.7 mTorr Ar atmosphere at room temperature, radio frequency power (15 W), sputtering time (60 s) Radio frequency sputtering: 13.56 MHz radio frequency sputtering, 4.7 mTorr Ar atmosphere at room temperature, radio frequency power (15 W), sputtering time (90 s) Radio frequency sputtering: 13.56 MHz radio frequency sputtering, 4.7 mTorr Ar atmosphere at room temperature, radio frequency power (10 W), sputtering time (30 s) Radio frequency sputtering: 13.56 MHz radio frequency sputtering, 4.7 mTorr Ar atmosphere at room temperature, radio frequency power (10 W), sputtering time (60 s) Radio frequency sputtering: 13.56 MHz radio frequency sputtering, 4.7 mTorr Ar atmosphere at room temperature, radio frequency power (10 W), sputtering time (120 s) Radio frequency sputtering: 13.56 MHz radio frequency sputtering, 4.7 mTorr Ar atmosphere at room temperature, radio frequency power (10 W), sputtering time (300 s) Radio frequency sputtering: 13.56 MHz radio frequency sputtering, 4.7 mTorr Ar atmosphere at room temperature, radio frequency power (10 W), sputtering time (600 s) Radio frequency sputtering: 13.56 MHz radio frequency sputtering, 0.003‒0.004 mTorr Ar atmosphere at room temperature, radio frequency power, silica target (250 W), radio frequency power, Cu target (25 W), cosputtering time (105 min) Spark plasma sintering: uniaxial pressure (50 MPa), in vacuum at 1020 °C for 3 min Staining process: pigment (red ochre); annealing (T < 660 °C for 1 h and 40 min) Staining process: pigment (red ochre); annealing (T < 660 °C for 1 h and 40 min) Ion-exchange technique GNCs prepared by staining process, spark plasma sintering, radio frequency sputtering, spray pyrolysis, and CVD techniques [371–376] Host matrix Composite Table 6 ‒ ‒ NLO devices β(720 nm) = 5.72 × 10−12 m·W−1 ‒ Art Application [374] [373] [373] [372] [371] Ref. Javier FONSECA. NPs embedded into glass matrices: GNCs 57 Ag NPs Ag NPs Ag NPs Ag NPs Mesoporous bioactive glassb) Mesoporous bioactive glassa) Mesoporous bioactive glassa) Mesoporous bioactive glassa) Silica glass Ag NPs@mesoporous bioactive glasses Ag NPs@Mesoporous bioactive glasses Ag NPs@mesoporous bioactive glasses Ag NPs@mesoporous bioactive glasses Au NPs@silica glass a) Silver acetate as precursor; b) silver nitrate as precursor. Ag NPs Mesoporous bioactive glassa) Ag NPs@mesoporous bioactive glasses Ag NPs Cu NPs Silica glass Cu NPs@silica glass Reinforcement Host matrix Composite (50) (9 ± 3) (8 ± 3) (8 ± 2) (7 ± 3) (9 ± 4) Size of NPs (min‒max (mean))/nm 12 ‒ 6.56 mol.% 3.28 mol.% 1.64 mol.% ‒ ‒ ‒ Loading of NPs (wt.% in excess) Radio frequency sputtering: 13.56 MHz radio frequency sputtering, 0.003‒0.004 mTorr Ar atmosphere at room temperature, radio frequency power, silica target (250 W), radio frequency power, Cu target (40 W), cosputtering time (90 min) Spray pyrolysis: dispersion of precursors into droplets, preheating at 400 °C, calcination at 700 °C, annealing 500 °C Spray pyrolysis: dispersion of precursors into droplets, preheating at 400 °C, calcination at 700 °C, annealing 500 °C Spray pyrolysis: dispersion of precursors into droplets, preheating at 400 °C, calcination at 700 °C, annealing 500 °C Spray pyrolysis: dispersion of precursors into droplets, preheating at 400 °C, calcination at 700 °C, annealing 500 °C Spray pyrolysis: dispersion of precursors into droplets, preheating at 400 °C, calcination at 700 °C, annealing 500 °C Chemical vapor deposition: aerosol assisted chemical vapor deposition, Au NPs in toluene, aerosol generation (ultrasonic humidification), N2 flow rate (1 L·min−1), glass temperature 450 °C Ion-exchange technique ‒ Antibacterial activity Antibacterial activity Application [376] [375] [154] Ref. (continued) 58 Front. Mater. Sci. 2022, 16(3): 220607 Javier FONSECA. NPs embedded into glass matrices: GNCs reduced to metal atoms and these grow and coalesce into metal NPs within the glass matrix. The metal NPs together with the pigment interact with light and, therefore, color the glass [377]. The staining process has been applied to an aluminosilicate glass (SiO2‒Al2O3‒Na2O‒CaO‒MgO) to study the formation of NPs, the evolution of the glass matrix and the final color of the prepared GNC [371]. The glass was enameled with a mixture of red ochre and Ag2SO4 (to obtain a yellow GNC) or a mixture of red ochre and Cu2SO4 (to obtain a red GNC). Subsequently, the glass was treated at a temperature up to 660 °C for 1 h and 40 min. This annealing favored the Ag+ (or Cu+)‒Na+ ion-exchange, the neutralization of those Ag+ (or Cu+) ions and, finally, the precipitation of Ag (or Cu) NPs within the matrix. It was suggested that Sn2+ ions can favor the precipitation of Ag (or Cu) NPs as follows [378]: Sn2+ → Sn4+ + 2e− (49) Ag+ + e− → Ag0 (50) Cu+ + e− → Cu0 (51) Cu2+ + 2e− → Cu0 (52) The size of the embedded Ag NPs was reported to range from 2.5 to 6 nm. The size of the Cu NPs ranged between 25 and 40 nm. These NPs were a core‒shell system where the core was Ag (or Cu) NPs, and the shell was probably an oxidized form of Ag (or Cu). The Ag NPs were homogeneously distributed on the glass surface. However, the Cu NPs were accumulated in Cu-rich regions. There was a decrease in Q4 species and an increased in Q2 and Q3 species due to the staining process. Therefore, this process caused the depolymerization of the glass surface as a classical ion-exchange process does. The depolymerization originated by ion-exchange: Na−O−Si≡ ↔ M−O−Si≡ 3.5.2 Spark plasma sintering Spark plasma sintering (SPS), also known as pulsed electric current sintering (PECS) [379], field-assisted sintering technique (FAST) [380‒381], plasma-activated sintering (PAS) [382] and current-activated pressureassisted densification (CAPAD) [383‒384], has gained importance in the last 20 years. SPS is a technique that allows materials to be densified at low temperatures in a short time [385‒386]. This technique is based on the simultaneous application of pulsed direct current, heat, and uniaxial pressure on a sample (as far as this review is concerned, a mixture of GNC precursors) loaded into a graphite die (Fig. 31) [387]. The process can be carried out in different environments such as vacuum, argon, hydrogen, air, etc. It is generally claimed that the high densification rates come primarily from the use of highenergy pulsed direct current. Pulsed direct current is often achieved by applying a voltage of about 30 V and a current of 600‒1000 A. The duration of each pulse varies between 1 and 300 ms [387]. The electrical discharges across the sample can generate microcosmic spark discharge and plasma. There is a debate about the existence of plasma and spark discharge [388‒389]. It is primarily believed that the spark discharges appear in the gaps among the sintered particles under appropriate conditions in SPS [390]. After the initial spark discharge stage, spark plasma would be generated [391]. The high local temperature, which can be generated momentarily in the initial spark discharge stage, can modify the GNC precursors. The total synthesis time depends on the material to be prepared [387]. Zhang et al. have embedded Au NPs within a silica glass matrix by applying SPS to SBA-15 with confined (53) where M is the metal ion. However, the glass surface was also affected by a flow of oxygen at the interface. ≡Si−O−Si≡ + 2M+ + (1/2)O2 ↔ 2(M−O−Si≡) (54) Therefore, ion-exchange and oxygen flow at the interface were responsible of the depolymerization [371]. 59 Fig. 31 Schematic illustration of the features of a spark plasma sintering apparatus. 60 Front. Mater. Sci. 2022, 16(3): 220607 Au NPs [372]. It is worth mentioning that SBA-15 is a well-ordered hexagonal mesoporous silica with uniform pore sizes [392]. First, Au NPs with a size of 3 to 10 nm were encapsulated within the pores of SBA-15 by postsynthetic functionalization. This encapsulation did not modify the ordered hexagonal mesostructure of SBA-15. Then, the composite (SBA-15 with confined Au NPs) was loaded into a graphite die and sintered using SPS under a uniaxial pressure of 50 MPa in vacuum at 1020 °C for 3 min. Thus, SBA-15 with confined Au NPs became glass silica with embedded Au NPs. The Au NPs were found to be homogenously distributed within the glass matrix. The size of these NPs also ranged from 3 to 10 nm. Therefore, the size of the NPs was preserved. At the wavelength of 720 nm, n2 and β of this GNC were 1.74 × 10−18 m2·W−1 and 5.72 × 10−12 m·W−1, respectively [372]. This method has also been applied to SBA-15 with confined Ag NPs or Pt NPs to obtain Ag NPs or Pt NPs embedded within silica glass [393]. 3.5.3 Radio frequency sputtering by sputtering; convection of the gas phase target to the substrate; and the condensation, nucleation and growth of the target on the substrate. Therefore, the process is a solid‒gas‒solid reaction [394‒395]. Gallium arsenide (GaAs) NPs and Au NPs have been embedded in silica glasses using radio frequency sputtering (13.56 MHz) [373]. Radio frequency sputtering was performed in a 4.7 mTorr Ar atmosphere at room temperature. The radio frequency power supplied to the GaAs and Au targets was 15 and 10 W, respectively. First, the GaAs target was sputtered onto a silica glass. To form GaAs NPs, the GaAs deposition time was shorter than the time required in forming a continuous layer. Then, SiO2 was sputtered onto the GaAs NPs supported on silica glass. Thus, the GaAs NPs were embedded within a silica glass. The average size of the GaAs NPs was 2.7, 4.4, and 7.7 nm for sputtering times of 30, 60, and 90 s, respectively (Fig. 33). Therefore, the size of the GaAs NPs increased as the sputtering time increased. The GaAs NPs that were close to each other coalesced as they grew. Most of the GaAs NPs were found to be connected to each other during a sputtering time of 240 s. The Au NPs were Radio frequency sputtering is a high-energy vaporization technique. It is based on generating an energetic wave through an inert gas in a vacuum chamber (Fig. 32). Thus, the inert gas ionizes. These high-energy ions bombard the target material, sputtering off the target atoms to inject them into a substrate. As far as this review is concerned, the substrate is a glass matrix. A surface atom of the target is only sputtered if it receives enough energy to exceed its binding energy. High sputtering rates are obtained for frequencies in the MHz region. In brief, the sputterdeposition process comprises vaporizing a target material Fig. 32 Schematic illustration of the radio frequency sputtering process. Fig. 33 TEM images of GaAs NPs embedded in silica glass prepared by radio frequency sputtering for (a) 30 s, (b) 60 s, (c) 90 s, (d) 120 s, and (e) 240 s. Reproduced with permission from Ref. [373] (Copyright 1997 AIP Publishing). Javier FONSECA. NPs embedded into glass matrices: GNCs similarly embedded in a silica glass. First, Au was deposited on the silica glass. Subsequently, SiO2 was sputtered to bury the Au NPs. The average diameter of Au NPs for sputtering times of 30, 60, 120, 300, and 600 s was 2.43, 3.64, 8.09, 17.5, and 47.08 nm, respectively (Fig. 34). The growth kinetics of GaAs and Au NPs were different. While the number of Au NPs was maintained as the sputtering time increased, the number of GaAs decreased under the same conditions. This was attributed to a larger mobility of the GaAs precursors and NPs than those of Au [373]. Cattaruzza et al. have embedded Co NPs in silica glass by 13.56 MHz radio frequency sputtering [374]. The depositions were performed in an atmosphere of Ar at a pressure ranging from 30 × 10−4 to 40 × 10−4 mbar and at room temperature. Silica and Co were co-sputtered on silica glass for 120, 105, and 90 min. The radio frequency power supplied to the silica target was 250 W. The radio frequency power supplied to the Co target was 10, 25, and 40 W when co-sputtering was 120, 105, and 90 min, respectively. After co-sputtering, silica was sputtered for 5 min. Both Co0 atoms and Co2+ ions were found within the silica glasses. Finally, the samples were treated to promote the precipitation of Co0 atoms to Co NPs within the glass matrix. Annealing was carried out in a H2 atmosphere at 900 °C for 2 h, in air at 700 °C for 5 h, or in air at 700 °C for 5 h followed by annealing in a H2 atmosphere at 900 °C for 2 h. Thermal annealing in a H2 atmosphere promoted the precipitation of Co0 atoms to form Co NPs. Thermal treatment in air favored the formation of Co2+ ions within the silica glass. Subsequent reducing treatment did not promote Co NPs formation. The average size of the Co NPs obtained after annealing in H2 atmosphere the samples prepared with a radio frequency power of 25 and 40 W was approximately 14 and 12 nm, respectively. These results were attributed to larger nucleation sites in the GNC with the higher concentration of Co. In conclusion, radio frequency cosputtering together with heat treatments allows achieving different GNCs [374]. This research group has also incorporated AuCu alloy NPs into silica glass using 13.56 MHz radio frequency co-sputtering [396‒397]. The depositions were carried out in an Ar atmosphere at a pressure of 35×10−2 Pa. Whereas the radio frequency power supplied to the silica target was set at 250 W, the powers supplied to the Au and Cu targets changed from 5 to 13 W and from 5 to 17 W, respectively. The radio frequency power was established to obtain Au:Cu ratios within the glass matrices close to 0, 1, 2, 3, and 5. Silica, Au and Cu were co-deposited on a silica glass for 50 min. After co-sputtering, silica was sputtered for 5 min. Finally, the composites were annealed in a reducing atmosphere at 900 °C for 2 h. The Cu atoms, which were mostly in oxidized forms, were reduced during annealing. After reduction of the Cu atoms, the already formed Aurich nanoclusters were converted into AuCu alloy NPs. Therefore, annealing favored the formation of AuCu alloy NPs of different composition, depending on the relative amount of Au and Cu within the glass matrix. The average size of the embedded alloy NPs, regardless of the alloy composition, was found to range from 4 to 10 nm (Fig. 35) [396‒397]. 3.5.4 Fig. 34 TEM images of Au NPs embedded in silica glass prepared by radio frequency sputtering for (a) 30 s, (b) 60 s, (c) 120 s, (d) 300 s, and (e) 600 s. Reproduced with permission from Ref. [373] (Copyright 1997 AIP Publishing). 61 Spray pyrolysis Spray pyrolysis (SP) is a versatile and flexible technique for producing materials with a wide range of composition, size, and morphology. SP includes all synthesis processes in which a solution is atomized and thermolyzed to obtain a suitable phase. The main mechanisms of particle 62 Front. Mater. Sci. 2022, 16(3): 220607 Fig. 35 Cross-sectional TEM images of the GNC made of Au47Cu53 alloy embedded within a glass matrix: (a) lowmagnification bright-field image (the inset is the sizehistogram); (b) high-resolution image of an Au47Cu53 alloy embedded within a glass matrix. Reproduced with permission from Ref. [397] (Copyright 2007 Elsevier). formation in SP are one-particle-per-drop and gas-toparticle conversion. In the later, the particles are created in the gas phase reactions. The reaction products can precipitate on the existing particle seeds producing nanostructured particles or nucleate to form new particles. In this review, we will focus on the one-particle-per-drop mechanism, which is the dominant formation procedure of GNCs by SP [398]. As mentioned above, SP can be applied to synthesize GNCs through the mechanism of one-particle-per-drop. Solutions, colloidal dispersions, emulsions, and sols can be used as precursors. These precursors are atomized by various techniques such as pneumatic (pressure, two-fluid, nebulizers), ultrasonic, and electrostatic approaches. For a specific atomizer, the physicochemical characteristics of the drops are determined by the density, viscosity, and surface tension of the precursor. Furthermore, the atomizers affect the rate of atomization, the drop size, and the drop velocity. The atomized precursors are driven into a series of reactors (Fig. 36). Many physical phenomena occur simultaneously in the diffusion drier (first step of SP): the solvent evaporates from the surface of the drop, the solvent vapors diffuse away from the drop into the gas phase, the solute diffuses towards the center of the drop, and the drop shrinks. It should be mentioned that the drops can coagulate before and during this first step. In doing so, two or more drops collide and coalesce into a larger drop. This can only happen if there is a liquid phase. In the calcination reactor (second step of SP), the precursor is completely decomposed and the solute precipitates. Therefore, the GNC is formed during this step. The particles of GNC are sintered and annealed in the annealing furnace (third and last step of SP) [399]. The production of GNCs with suitable characteristics requires control of the atomization, coagulation, evaporation, precipitation, decomposition, and sintering processes. The calcination atmosphere and the precursors also affect the chemistry of the GNC. Precursors with high solubility precipitate homogenously and tend to form solid spherical particles in a process named volume precipitation. In contrast, precursors with low solubility precipitate earlier at the surface rather than at the center of the particle. This is known as surface precipitation. The scaling up of SP processes is limited by the drop size and the coagulation process. Drops smaller than 5 µm in diameter are required to achieve an optimal production rate. In addition, drop coagulation should be avoided. Therefore, further developments in atomization, and coagulation control are needed to commercialize SP processes. Insufficient information on the fundamental characteristics of the precursor systems also limited the industrial applicability of this method. However, SP offers the opportunity to produce new materials. Shih et al. have developed GNCs by SP [154]. The composites were mesoporous bioactive glasses (MBGs) with embedded Ag NPs. The precursors of the MBGs were tetraethyl orthosilicate (TEOS), calcium nitrate (Ca(NO3)2), and triethyl phosphate (TEP). A solution of MBG precursors and surfactant (Pluronic F-127) in ethanol was mixed with a solution of Ag precursors (silver acetate (AgA) or silver nitrate (AgN) in DI water). The mixture was dispersed into fine droplets using an ultrasonic nebulizer. The droplets were preheated to evaporate the solvent, calcined to precipitate the solute and decompose the precursors, and annealed, at 400, 700, Javier FONSECA. NPs embedded into glass matrices: GNCs Fig. 36 63 Schematic illustration of the steps to prepare GNCs by spray pyrolysis. and 500 °C, respectively. The particle size of pure MBG, GNC prepared using AgA as a precursor, and GNC prepared using AgN as a precursor was (536 ± 174), (500 ± 158), and (494 ± 166) nm, respectively. Ag NPs were found on the surface of the composite prepared from AgA. The composite prepared with the AgN precursor contained the Ag NPs homogenously distributed within the glass matrix. It should be mentioned that AgN has greater solubility in water than AgA, which was suggested to facilitate the distribution of AgN before calcination. The average size of Ag NPs was (9 ± 4) and (7 ± 3) nm for the GNC prepared from AgA, and the GNC prepared from AgN, respectively. It was suggested that the Ag NPs grew larger in the composite prepared from AgA than in the composite prepared from AgN due to the higher relative concentration of Ag on the surface of the former. The BET surface area was reported to be (116 ± 1), (148 ± 6), and (202 ± 3) m2·g−1 for pure MBG, GNC prepared from AgA, and GNC prepared from AgN, respectively. Therefore, the incorporation of Ag NPs increased the surface area of the glass matrix. The GNCs showed bioactivity, which is defined as the ability to form hydroxyapatite (HA) layers after immersion in human body fluid. Furthermore, both composites exhibited antibacterial activity whereas pure MBG did not [154]. To determine the role of Ag NPs on the bioactivity of a MBG, the same research group has incorporated different concentrations (1.64, 3.28, and 6.56 mol.%) of this reinforcement in the glass using the SP technique [375]. The glass precursors were TEOS, Ca(NO3)2, TEP and a surfactant (Pluronic F-127). A solution of these precursors was mixed with a specific amount of AgA. This new mixture was dispersed into fine droplets with a nebulizer. The droplets were preheated, calcined, and annealed at 400, 700, and 500 °C, respectively. In other words, they were converted into GNC particles via SP. The size of the Ag NPs embedded in a MBG was found to be (8 ± 2), (8 ± 3), and (9 ± 3) nm for those GNCs that contained 1.64, 3.28, and 6.56 mol.% Ag, respectively. In these composites, most of the Ag NPs were located on the surface of the MBG. The BET surface area was (192.6 ± 0.4), (174.6 ± 10.5), (164.6 ± 22.1), and (148.2 ± 6.4) m2·g−1 for the MBG and the GNCs that contained 1.64, 3.28, and 6.56 mol.% Ag, respectively. While MBG did not show inhibition of bacterial growth, the antibacterial activity of the GNCs was found to increase as the concentration of embedded Ag NPs increased. Moreover, the Ag NPs also enhanced bioactivity. It should be mentioned that the Ag NPs embedded in a glass decrease the bioactivity of such glass if the reinforcement is located within the glass matrix [400]. Therefore, it was concluded that the influence of Ag NPs on the bioactivity of MBG depends on the location of these NPs (within the glass structure or on the glass surface) [375]. 3.5.5 Chemical vapor deposition Chemical vapor deposition (CVD) is a well-studied 64 Front. Mater. Sci. 2022, 16(3): 220607 technique for producing high quality 2D films on substrate surfaces at a reasonable cost [401‒403]. A film structure is not favorable when the attractive forces between the deposited materials are stronger than the forces between the substrate and the deposited materials. Under these conditions, the growth of the deposited material follows an island formation scheme [404‒405]. Moreover, the substrate surface can be functionalized to create areas with and without nucleation sites. Therefore, NPs can be controllably grown on substrates using the CVD technique [406]. The conversion of precursors into the desired products by CVD involves thermal decomposition, transport, and deposition [402]. A carrier gas facilitates the introduction of chemical precursor gas or gases into a reaction system containing a heated substrate to be coated. The deposition occurs on and near a hot surface substrate. The flow rate and partial pressure of each precursor can be varied to control the deposition. CVD can be accompanied by the production of chemical by-products. These are expelled from the chamber together with the unreacted precursor gases. High purity precursors are often required to avoid unwanted side reactions and by-products [407]. Ertorer et al. have embedded Au NPs in BK7 glass by CVD [406]. The BK7 glass was silanized with hexamethyldisilazane. This functionalization was carried out to create nucleation sites. An organometallic Au precursor ((trimethylphospine)methylgold = [(CH3)3P]AuCH3) was used in this CVD. The deposition of [(CH3)3P]AuCH3 on the functionalized substrate was performed in an Ar atmosphere at 65 °C and 5 Pa for 13, 15, 18, and 23 min. The −NH-terminated substrate surface was assumed to interact with the Au in the precursor. The phosphorous group of the precursor (−P−(CH3)3) was proposed to be displaced by the nitrogen group (−NH) on the substrate surface. Thus, Au was immobilized on the glass surface and the phosphorous group was released. Then, the Au from a second precursor molecule bound to the already immobilized gold species. The average size of the Au NPs Fig. 37 was 11 and 17 nm when the CVD time was 13 and 23 min, respectively. Therefore, the size of the Au NPs increased as the CVD time increased. Furthermore, increasing CVD time also resulted in NPs with less uniform size distribution and different morphologies, such as spherical, star-shaped, etc. For short CVD times, the only morphology of the Au NPs was spherical. These GNCs (Au NPs@BK7 glass) were found to be suitable for biomolecule-sensing applications [406]. As in the above example, conventional thermal CVD requires volatile precursors. Organometallic precursors are effective. However, they have several drawbacks: (1) they are often sensitive to air and humidity, (2) their synthesis can be difficult, and (3) they can incorporate impurities into the final composite. Inorganic precursors are generally more stable and give a cleaner decomposition. However, they are normally less volatile and require high temperatures and low pressures [408]. Aerosol-assisted chemical vapor deposition (AACVD) is a variant of the CVD process. AACVD uses liquid‒gas aerosols to transport soluble precursors to a heated substrate (Fig. 37). Therefore, the limitations of thermal stability and volatility are eliminated by designing precursors specifically for AACVD [409]. Au NPs have been deposited on silica glass using AACVD [376]. First, Au colloids were synthesized in toluene. In this solution, the average size of the Au NPs was 10 nm. Then, an aerosol was generated from this solution using an ultrasonic humidifier with an operating frequency of 40 kHz. The aerosol was transported to a horizontal-bed CVD reactor by N2 gas. The gas flow was continued until all the precursors passed through the reactor. The deposition was carried out at a glass temperature of 450 °C. The size of the Au NPs deposited on silica glass was around 50 nm in diameter. The Au NPs in the GNC were larger than those in the precursor solution. Therefore, the Au NPs agglomerated in the gas phase or on the surface of the substrate. Gas-phase particles are subject to a thermophoretic force when Schematic illustration of the AACVD system. Javier FONSECA. NPs embedded into glass matrices: GNCs exposed to a temperature gradient [410]. This force is directed away from the hot surface. Larger particles are repelled by a hot surface and attracted to a cold surface. Thus, the larger particles cannot diffuse through the thermal boundary layer on the surface of the hot substrate. Thermophoresis was the dominant force determining the location of Au NPs deposition. Therefore, the Au NPs were obtained exclusively on the substrate surface [376]. Although this is beyond the scope of this review, it is worth mentioning that the AACVD technique has been used extensively to deposit nanocomposite films on substrates [411‒413]. 4 Conclusions Research on GNCs has been very active and productive in the past decades. GNCs have played — and still do — an important role in the fields of optoelectronics, photonics, sensing, electrochemistry, catalysis, biomedicine, and art. Regarding optoelectronics, all-optical switching has been extensively investigated for decades because it can potentially overcome the speed limitation of electrical switching devices [30,414‒416]. However, all-optical switching devices require a large driving energy, which is a fundamental limitation for most applications. There is a trade-off between speed and energy. In other words, energy can be reduced by sacrificing speed. The large energy required is attributed to the inherently small optical non-linearity in existing materials. Therefore, future opportunities in all-optical switching will depend on the preparation of materials with enhanced Kerr susceptibilities, as these materials are still at a relatively early stage of development. Particularly, GNCs are expected to break this trade-off. In the field of photonics, the addition of noble metal NPs to REs-doped glasses further enhances their efficiency as solid-state light sources. Noble metal NPs improve the emission intensity of RE ions. These GNCs are also promising photovoltaic (PV) cells. Silicon-based PV systems cannot efficiently convert a large part of the solar radiation in the UV spectrum into electricity. However, GNCs can overcome this limitation. GNCs have potential as low-cost photonic systems with improved characteristics, but further research is required. GNCs can be ideal SERS platforms. Low-cost sensors and biosensors with ultra-sensitive detection limits can be made of GNCs [360,364,417‒418]. They can provide high 65 efficiency (in terms of signal enhancement) and reuse. However, NPs are prone to oxidation and sulfidation phenomena. Therefore, ensuring the long-term stability of metal NPs within glass matrices is an indispensable condition for developing GNCs-based sensors and biosensors. The preparation of GNCs as SERS substrates remains a challenging task. Extensive work and research have been done to develop solid materials capable of replacing liquid electrolytes. SEs can improve the performance, safety, and durability of batteries [419]. The performance of SSBs is highly dependent on ion diffusion within the electrolyte. SEs must exhibit high ionic conductivity and very low electronic conductivity [420]. An ideal SE is a pure ion conductor. Thermal and chemical stabilities are also important factors of SEs [421]. Therefore, GNCs made of extremely high concentrations of superionic metastable NPs embedded within stable glasses are promising candidates for SEs applicable to SSBs [422]. Inexpensive, environmentally friendly, durable, reusable, and inert glasses are attractive supports for catalytically active metal NPs. The main drawback of these glasses is their lack of porosity. Therefore, further research is expected for the development of porous glasses. GNCs made of porous glasses and metal NPs can become competitive catalysts. Recent advances in the fields of bone regeneration and cancer therapy focus on the development of MBGs [423‒424]. These glasses also have potential for targeted drug delivery [425]. The incorporation of NPs within MBGs can improve the mechanical and antibacterial properties and the biodegradation rates of the pristine MBGs [426]. However, many questions about the usability of MBG-based GNCs for biomedicine have yet to be addressed: optimal dose of NPs and drugs to incorporate into MBG, fate of MBG in vivo, long-term side effects, etc. The conservation–restoration of GNCs-based Historical Heritage consists of preventing their deterioration and applying the appropriate treatments to restore the masterpieces to their original state without permanently altering them. Best practices employ the protocol of “avoiding, blocking, detecting, and responding” to any threat. It is also important to assess risks and needs based on the main agents of deterioration. Cleaning methods have evolved from washing with water to using special chemical cleaners made especially for GNCs. In addition, cleaning methods also differ if the GNCs are already 66 Front. Mater. Sci. 2022, 16(3): 220607 damaged, fragile, or very old. It should be noted that today the conservation-restoration of the GNCs-based Historical Heritage is strongly promoted. The most important synthesis methods of GNCs (meltquenching, sol-gel, ion implantation, ion-exchange, staining process, spark plasma sintering, radio-frequency sputtering, spray pyrolysis, and CVD techniques) are extensively described in this review. Considering that future directions of GNC synthesis appear to lie in the combination of various GNC synthesis approaches, such as ion implantation and radio frequency sputtering or laser irradiation [35], a thorough understanding of GNC synthesis methods is the first step towards unprecedented GNCs development. Furthermore, in-depth knowledge of GNC synthesis techniques may facilitate innovative uses of such techniques [427]. In this review, we have also explored the growth of NPs within glass matrices. The size-controlled preparation of NPs within glass matrices, which remains a challenge, is essential for advanced applications. The field of GNCs will continue to provide opportunities for those seeking new engineering materials. In addition, despite all the advances in our knowledge about GNCs, there are still fundamental questions that deserve to be investigated. For instance, the relationships between the chemical composition, structure, and properties of GNCs are still far from scientific understanding. There is no doubt that the field of GNCs will continue to grow and flourish. These are exciting times ahead. Acknowledgements The Chemical Engineering Northeastern University supported this work. Department at [5] Shang L, Bian T, Zhang B, et al. Graphene-supported ultrafine metal nanoparticles encapsulated by mesoporous silica: robust catalysts for oxidation and reduction reactions. Angewandte Chemie International Edition, 2014, 53(1): 250–254 [6] Yuan Q, Duan H H, Li L L, et al. Homogeneously dispersed ceria nanocatalyst stabilized with ordered mesoporous alumina. Advanced Materials, 2010, 22(13): 1475–1478 [7] Ohko Y, Tatsuma T, Fujii T, et al. Multicolour photochromism of TiO2 films loaded with silver nanoparticles. Nature Materials, 2003, 2(1): 29–31 [8] Lykhach Y, Staudt T, Tsud N, et al. Enhanced reactivity of Pt nanoparticles supported on ceria thin films during ethylene dehydrogenation. Physical Chemistry Chemical Physics, 2011, 13(1): 253–261 [9] Mitra A, De G. Chapter 6: Sol-gel synthesis of metal nanoparticle incorporated oxide films on glass. Glass Nanocomposites: Synthesis, Properties and Applications, 2016: 145–163 [10] Fonseca J, Lu J. Single-atom catalysts designed and prepared by the atomic layer deposition technique. ACS Catalysis, 2021, 11(12): 7018–7059 [11] Karmakar B. Chapter 1: Fundamentals of glass and glass nanocomposites. Glass Nanocomposites: Synthesis, Properties and Applications, 2016: 3–53 [12] Zanotto E D, Mauro J C. The glassy state of matter: its definition and ultimate fate. Journal of Non-Crystalline Solids, 2017, 471: 490–495 [13] Fonseca J, Gong T, Jiao L, et al. Metal-organic frameworks (MOFs) beyond crystallinity: amorphous MOFs, MOF liquids and MOF glasses. Journal of Materials Chemistry A, 2021, 9(17): 10562–10611 [14] Xiang W, Gao H, Ma L, et al. Valence state control and third- References order nonlinear optical properties of copper embedded in sodium borosilicate glass. ACS Applied Materials & Interfaces, 2015, 7(19): 10162–10168 [1] Lee J, Mahendra S, Alvarez P J J. Nanomaterials in the [15] Zhong J, Xiang W, Chen Z, et al. Microstructures and third- construction industry: a review of their applications and order optical nonlinearities of Cu2In nanoparticles in glass environmental health and safety considerations. ACS Nano, matrix. Journal of Alloys and Compounds, 2013, 572: 137–144 2010, 4(7): 3580–3590 [2] Yang Z, Ren J, Zhang Z, et al. Recent advancement of nanostructured carbon for energy applications. Chemical Reviews, 2015, 115(11): 5159–5223 [3] Aricò A S, Bruce P, Scrosati B, et al. Nanostructured materials for advanced energy conversion and storage devices. Nature Materials, 2005, 4(5): 366–377 [4] Sanchez F, Sobolev K. Nanotechnology in concrete — a review. Construction and Building Materials, 2010, 24(11): 2060–2071 [16] Chatterjee S, Saha S K, Chakravorty D. Chapter 2: Glass-based nanocomposites. Glass Nanocomposites: Synthesis, Properties and Applications, 2016: 57–88 [17] Amendola V, Pilot R, Frasconi M, et al. Surface plasmon resonance in gold nanoparticles: a review. Journal of Physics: Condensed Matter, 2017, 29(20): 203002 [18] Korgel B A. Materials science: composite for smarter windows. Nature, 2013, 500(7462): 278–279 [19] Mangin S, Gottwald M, Lambert C H, et al. Engineered materials for all-optical helicity-dependent magnetic switching. Javier FONSECA. NPs embedded into glass matrices: GNCs [34] Chakraborty P. Metal nanoclusters in glasses as non-linear Nature Materials, 2014, 13(3): 286–292 [20] Llordés A, Garcia G, Gazquez J, et al. Tunable near-infrared and visible-light transmittance 67 in nanocrystal-in-glass photonic materials. Journal of Materials Science, 1998, 33(9): 2235–2249 [35] Haglund R F Jr. Ion implantation as a tool in the synthesis of composites. Nature, 2013, 500(7462): 323–326 [21] Castro J M, Geraghty D F, West B R, et al. Fabrication and comprehensive modeling of ion-exchanged Bragg optical adddrop multiplexers. Applied Optics, 2004, 43(33): 6166–6173 practical third-order nonlinear optical materials. Materials Science and Engineering A, 1998, 253(1‒2): 275–283 [36] Weber M J, Milam D, Smith W L. Nonlinear refractive index of [22] Veasey D L, Funk D S, Peters P M, et al. Yb/Er-codoped and glasses and crystals. Optical Engineering, 1978, 17(5): 463–469 Yb-doped waveguide lasers in phosphate glass. Journal of Non- [37] Ryasnyansky A I, Palpant B, Debrus S, et al. Nonlinear optical Crystalline Solids, 2000, 263–264: 369–381 properties of copper nanoparticles synthesized in indium tin [23] Tervonen A, West B R, Honkanen S K. Ion-exchanged glass waveguide technology: a review. Optical Engineering, 2011, 50(7): 071107 oxide matrix by Ion implantation. Journal of the Optical Society of America B, 2006, 23(7): 1348–1353 [38] Chemla D S, Miller D A B. Room-temperature excitonic [24] Choi J, Bellec M, Royon A, et al. Three-dimensional direct nonlinear-optical effects in semiconductor quantum-well femtosecond laser writing of second-order nonlinearities in structures. Journal of the Optical Society of America B, 1985, glass. Optics Letters, 2012, 37(6): 1029–1031 2(7): 1155–1173 [25] Royon A, Bourhis K, Bellec M, et al. Silver clusters embedded [39] Kawabata A, Kubo R. Electronic properties of fine metallic in glass as a perennial high capacity optical recording medium. particles. II. Plasma resonance absorption. Journal of the Advanced Materials, 2010, 22(46): 5282–5286 Physical Society of Japan, 1966, 21(9): 1765–1772 [26] Azlan M N, Hajer S S, Halimah M K, et al. Comprehensive comparison on optical properties of samarium oxide [40] Halperin W P. Quantum size effects in metal particles. Reviews of Modern Physics, 1986, 58(3): 533–606 (micro/nano) particles doped tellurite glass for optoelectronics [41] Haglund R F Jr, Yang L, Magruder R H III, et al. Nonlinear applications. Journal of Materials Science Materials in optical properties of metal-quantum-dot composites synthesized Electronics, 2021, 32(11): 14174–14185 by ion implantation. Nuclear Instruments and Methods in [27] Oh Y J, Jeong K H. Glass nanopillar arrays with nanogap-rich silver nanoislands for highly intense surface enhanced Raman scattering. Advanced Materials, 2012, 24(17): 2234–2237 [28] Kurobori T, Nakamura S. A novel disk-type X-ray area imaging detector using radiophotoluminescence in silveractivated phosphate glass. Radiation Measurements, 2012, 47(10): 1009–1013 patient during percutaneous coronary intervention procedures using radiophotoluminescence glass dosimeters. Radiation Protection Dosimetry, 2017, 175(1): 31–37 [30] Werner S, Navaridas J, Luján M. A survey on optical networkon-chip architectures. ACM Computing Surveys, 2017, 50(6): 1–37 [31] Heidt A M, Li Z, Sahu J, et al. 100 kW Peak power picosecond thulium-doped fiber amplifier system seeded by a gainswitched diode laser at 2 µm. Optics Letters, 2013, 38(10): 1615–1617 [32] Li M, Huang S, Wang Q, et al. Nonlinear lightwave circuits in chalcogenide glasses fabricated by ultrafast laser. Optics Letters, 2014, 39(3): 693–696 [33] Najafi S I, Lefebvre P, Albert J, et al. Ion-exchanged Mach‒Zehnder interferometers in glass. Applied Optics, 1992, 31(18): 3381–3383 and Atoms, 1994, 91(1‒4): 493–504 [42] Garland J C, Tanner D B. Electrical Transport and Optical Properties of Inhomogeneous Media. New York, USA: American Institute of Physics, 1978 [43] Li Y Q, Sung C C, Inguva R, et al. Nonlinear-optical properties of semiconductor composite materials. Journal of the Optical [29] Kato M, Chida K, Moritake T, et al. Direct dose measurement on Physics Research Section B: Beam Interactions with Materials Society of America B, 1989, 6(4): 814–817 [44] Hale D K. The physical properties of composite materials. Journal of Materials Science, 1976, 11(11): 2105–2141 [45] Haus J W, Kalyaniwalla N, Inguva R, et al. Nonlinear-optical properties of conductive spheroidal particle composites. Journal of the Optical Society of America B, 1989, 6(4): 797–807 [46] Hou W, Cronin S B. A review of surface plasmon resonanceenhanced photocatalysis. Advanced Functional Materials, 2013, 23(13): 1612–1619 [47] Kabi S, Ghosh A. Structural investigation on silver phosphate glasses embedded with nanoparticles. Journal of Alloys and Compounds, 2012, 520: 238–243 [48] Wang Q Q, Han J B, Gong H M, et al. Linear and nonlinear optical properties of Ag nanowire polarizing glass. Advanced Functional Materials, 2006, 16(18): 2405–2408 [49] Lin G, Pan H H, Qiu J R, et al. Nonlinear optical properties of lead nanocrystals embedding glass induced by thermal 68 Front. Mater. Sci. 2022, 16(3): 220607 treatment and femtosecond laser irradiation. Chemical Physics Letters, 2011, 516(4‒6): 186–191 3427–3430 [63] Ghosh S K, Nath S, Kundu S, et al. Solvent and ligand effects [50] Qu S, Zhang Y, Li H, et al. Nanosecond nonlinear absorption in on the localized surface plasmon resonance (LSPR) of gold Au and Ag nanoparticles precipitated glasses induced by a colloids. The Journal of Physical Chemistry B, 2004, 108(37): femtosecond laser. Optical Materials, 2006, 28(3): 259–265 13963–13971 [51] Lin G, Tan D Z, Luo F F, et al. Linear and nonlinear optical [64] Rechberger W, Hohenau A, Leitner A, et al. Optical properties properties of glasses doped with Bi nanoparticles. Journal of of two interacting gold nanoparticles. Optics Communications, Non-Crystalline Solids, 2011, 357(11‒13): 2312–2315 2003, 220(1‒3): 137–141 [52] Hache F, Ricard D, Flytzanis C. Optical nonlinearities of small [65] Jain P K, Qian W, El-Sayed M A. Ultrafast electron relaxation metal particles: surface-mediated resonance and quantum size dynamics in coupled metal nanoparticles in aggregates. The effects. Journal of the Optical Society of America B, 1986, 3(12): 1647–1655 [53] Christensen N E, Seraphin B O. Relativistic band calculation and the optical properties of gold. Physical Review B, 1971, 4(10): 3321–3344 Journal of Physical Chemistry B, 2006, 110(1): 136–142 [66] Su K H, Wei Q H, Zhang X, et al. Interparticle coupling effects on plasmon resonances of nanogold particles. Nano Letters, 2003, 3(8): 1087–1090 [67] Jain P K, Eustis S, El-Sayed M A. Plasmon coupling in nanorod [54] Hache F, Ricard D, Flytzanis C, et al. The optical kerr effect in assemblies: optical absorption, discrete dipole approximation small metal particles and metal colloids: The case of gold. simulation, and exciton-coupling model. The Journal of Applied Physics A: Materials Science & Processing, 1988, 47: Physical Chemistry B, 2006, 110(37): 18243–18253 347–357 [68] De S, De G. In-situ generation of Au nanoparticles in UV- [55] Liz-Marzán L M. Tailoring surface plasmons through the curable refractive index controlled SiO2‒TiO2‒PEO hybrid morphology and assembly of metal nanoparticles. Langmuir, films. The Journal of Physical Chemistry C, 2008, 112(28): 2006, 22(1): 32–41 10378–10384 [56] El-Sayed M A. Some interesting properties of metals confined [69] Medda S K, Mitra M, De G. Tuning of Ag-SPR band position in time and nanometer space of different shapes. Accounts of in refractive index controlled inorganic–organic hybrid Chemical Research, 2001, 34(4): 257–264 SiO2–PEO–TiO2 films. Journal of Chemical Sciences, 2008, [57] Link S, El-Sayed M A. Shape and size dependence of radiative, 120(6): 565–572 non-radiative and photothermal properties of gold nanocrystals. [70] Zhang J, Fu Y, Chowdhury M H, et al. Plasmon-coupled International Reviews in Physical Chemistry, 2000, 19(3): fluorescence probes: effect of emission wavelength on 409–453 fluorophore-labeled silver particles. The Journal of Physical [58] Moores A, Goettmann F. The plasmon band in noble metal nanoparticles: an introduction to theory and applications. New Journal of Chemistry, 2006, 30(8): 1121–1132 [59] Jain P K, Lee K S, El-Sayed I H, et al. Calculated absorption Chemistry C, 2008, 112(25): 9172–9180 [71] Zhang Y, Dragan A, Geddes C D. Wavelength dependence of metal-enhanced fluorescence. The Journal of Physical Chemistry C, 2009, 113(28): 12095–12100 and scattering properties of gold nanoparticles of different size, [72] Zhang J, Fu Y, Chowdhury M H, et al. Metal-enhanced single- shape, and composition: applications in biological imaging and molecule fluorescence on silver particle monomer and dimer: biomedicine. The Journal of Physical Chemistry B, 2006, coupling effect between metal particles. Nano Letters, 2007, 110(14): 7238–7248 7(7): 2101–2107 [60] Eustis S, El-Sayed M A. Why gold nanoparticles are more [73] Katsuaki T. Field enhancement around metal nanoparticles and precious than pretty gold: noble metal surface plasmon nanoshells: a systematic investigation. The Journal of Physical resonance and its enhancement of the radiative and nonradiative Chemistry C, 2008, 112(40): 15721–15728 properties of nanocrystals of different shapes. Chemical Society Reviews, 2006, 35(3): 209–217 [61] De G, Medda S K, De S, et al. Metal nanoparticle doped [74] Zhu J. Spatial dependence of the local field enhancement in dielectric shell coated silver nanospheres. Applied Surface Science, 2007, 253(21): 8729–8733 coloured coatings on glasses and plastics through tuning of [75] Jupri S A, Ghoshal S K, Yusof N N, et al. Influence of surface surface plasmon band position. Bulletin of Materials Science, plasmon resonance of Ag nanoparticles on photoluminescence 2008, 31(3): 479–485 of Ho3+ ions in magnesium‒zinc-sulfophosphate glass system. [62] Underwood S, Mulvaney P. Effect of the solution refractive index on the color of gold colloids. Langmuir, 1994, 10(10): Optics and Laser Technology, 2020, 126: 106134 [76] Sontakke A D, Biswas K, Annapurna K. Concentration- Javier FONSECA. NPs embedded into glass matrices: GNCs dependent luminescence of Tb3+ ions in high calcium aluminosilicate glasses. Journal of Luminescence, 2009, 69 Luminescence, 2017, 188: 172–179 [89] Zmojda J, Miluski P, Kochanowicz M. Nanocomposite antimony-germanate-borate glass fibers doped with Eu3+ ions 129(11): 1347–1355 [77] Karunakaran R T, Marimuthu K, Surendra Babu S, et al. Structural, optical and thermal studies of Eu3+ ions in lithium fluoroborate glasses. Solid State Sciences, 2009, 11(11): with self-assembling silver nanoparticles for photonic applications. Applied Sciences, 2018, 8(5): 790 [90] Fares H, Elhouichet H, Gelloz B, et al. Surface plasmon resonance induced Er3+ photoluminescence enhancement in 1882–1889 [78] Nageswara Raju C, Sailaja S, Hemasundara Raju S, et al. Emission analysis of CdO–Bi2O3–B2O3 glasses doped with tellurite glass. Journal of Applied Physics, 2015, 117(19): 193102 Eu3+ and Tb3+. Ceramics International, 2014, 40(6): 7701–7709 [91] Kichanov S E, Kozlenko D P, Gorshkova Y E, et al. Structural [79] Vařák P, Nekvindová P, Vytykáčová S, et al. Near-infrared studies of nanoparticles doped with rare-earth ions in photoluminescence enhancement and radiative energy transfer oxyfluoride lead-silicate glasses. Journal of Nanoparticle in RE-doped zinc-silicate glass (RE = Ho, Er, Tm) after silver ion exchange. Journal of Non-Crystalline Solids, 2021, 557: Research, 2018, 20(3): 54 [92] Moskovits M. Surface-enhanced spectroscopy. Reviews of Modern Physics, 1985, 57(3): 783–826 120580 [80] Canevali C, Mattoni M, Morazzoni F, et al. Stability of [93] Nie S, Emory S R. Probing single molecules and single luminescent trivalent cerium in silica host glasses modified by nanoparticles by surface-enhanced Raman scattering. Science, boron and phosphorus. Journal of the American Chemical 1997, 275(5303): 1102–1106 [94] Simo A, Joseph V, Fenger R, et al. Long-term stable silver Society, 2005, 127(42): 14681–14691 [81] Cao W, Huang F, Wang Z, et al. Controllable structural Er3+-doped tailoring for enhanced luminescence in highly germanosilicate glasses. Optics Letters, 2018, 43(14): subsurface ion-exchanged glasses for SERS applications. ChemPhysChem, 2011, 12(9): 1683–1688 [95] Haynes C L, McFarland A D, Van Duyne R P. Surfaceenhanced Raman spectroscopy. Analytical Chemistry, 2005, 3281–3284 [82] Huang F, E F, Lei R, et al. Enhancing the Er3+: Infrared and mid-infrared emission performance in germanosilicate-zinc 77(17): 338A–346A [96] Schatz G C. Theoretical studies of surface enhanced Raman scattering. Accounts of Chemical Research, 1984, 17(10): glasses. Journal of Luminescence, 2019, 213: 370–375 [83] Vijayakumar R, Marimuthu K. Luminescence studies on Ag nanoparticles embedded Eu3+ doped boro-phosphate glasses. 370–376 [97] Dousti M R, Sahar M R, Amjad R J, et al. Surface enhanced Raman scattering and up-conversion emission by silver Journal of Alloys and Compounds, 2016, 665: 294–303 [84] Kamrádek M, Kašík I, Aubrecht J, et al. Nanoparticle and solution doping for efficient holmium fiber lasers. IEEE nanoparticles in erbium-zinc-tellurite glass. Journal of Luminescence, 2013, 143: 368–373 [98] Moskovits M. Surface-enhanced Raman spectroscopy: a brief Photonics Journal, 2019, 11(5): 1–10 [85] Baker C C, Friebele E J, Burdett A A, et al. Nanoparticle perspective. In: Kneipp K, Moskovits M, Kneipp H, eds. doping for high power fiber lasers at eye-safer wavelengths. Surface-enhanced Raman scattering. Topics in Applied Physics. Berlin, Heidelberg: Springer, 2006: 103 Optics Express, 2017, 25(12): 13903–13915 [86] Zhang W, Lin J, Cheng M, et al. Radiative transition, local field enhancement and energy transfer microcosmic mechanism of tellurite glasses containing Er3+, Yb3+ ions and Ag [99] Campion A, Kambhampati P. Surface-enhanced Raman scattering. Chemical Society Reviews, 1998, 27(4): 241–250 [100] Qian X M, Nie S M. Single-molecule and single-nanoparticle nanoparticles. Journal of Quantitative Spectroscopy and SERS: Radiative Transfer, 2015, 159: 39–52 applications. Chemical Society Reviews, 2008, 37(5): 912–920 from fundamental mechanisms to biomedical [87] Meng S, Zhao G, Hou J, et al. High performance of near- [101] Fleischmann M, Hendra P J, McQuillan A J. Raman spectra of infrared emission for S-band amplifier from Tm3+-doped pyridine adsorbed at a silver electrode. Chemical Physics bismuth glass incorporated with Ag nanoparticles. Journal of Letters, 1974, 26(2): 163–166 [102] Liu D, Chen X, Hu Y, et al. Raman enhancement on ultra-clean Luminescence, 2020, 224: 117313 [88] Sgibnev Y M, Nikonorov N V, Ignatiev A I. High efficient graphene quantum dots produced by quasi-equilibrium plasma- luminescence of silver clusters in ion-exchanged antimony- enhanced chemical vapor deposition. Nature Communications, doped photo-thermo-refractive glasses: influence of antimony content and heat treatment parameters. Journal of 2018, 9: 193 [103] Kneipp J, Kneipp H, Kneipp K. SERS — a single-molecule and 70 Front. Mater. Sci. 2022, 16(3): 220607 nanoscale tool for bioanalytics. Chemical Society Reviews, electrolytes for lithium batteries. Polymer, 2006, 47(16): 5952–5964 2008, 37(5): 1052–1060 [104] Wells S M, Retterer S D, Oran J M, et al. Controllable nanofabrication of aggregate-like nanoparticle substrates and evaluation for surface-enhanced Raman spectroscopy. ACS [118] Munshi M Z A. Handbook of Solid State Batteries and Capacitors. Singapore: World Scientific, 1995 [119] Troy S, Schreiber A, Reppert T, et al. Life cycle assessment and resource analysis of all-solid-state batteries. Applied Nano, 2009, 3(12): 3845–3853 [105] Wang H, Levin C S, Halas N J. Nanosphere arrays with Energy, 2016, 169: 757–767 raman [120] Mohamed E A, Nabhan E, Ratep A, et al. Influence of BaTiO3 spectroscopy substrates. Journal of the American Chemical nanoparticles/clusters on the structural and dielectric properties Society, 2005, 127(43): 14992–14993 of glasses-nanocomposites. Physical Review B: Condensed controlled sub-10-nm gaps as surface-enhanced [106] Pal P, Bonyár A, Veres M, et al. A generalized exponential Matter, 2020, 589: 412220 relationship between the surface-enhanced Raman scattering [121] Mohamed E A, Moustafa M G, Kashif I. Microstructure, (SERS) efficiency of gold/silver nanoisland arrangements and thermal, optical and dielectric properties of new glass their non-dimensional interparticle distance/particle diameter nanocomposites of SrTiO3 nanoparticles/clusters in tellurite ratio. Sensors and Actuators A: Physical, 2020, 314: 112225 glass matrix. Journal of Non-Crystalline Solids, 2018, 482: [107] Chou C M, Thanh Thi L T, Quynh Nhu N T, et al. Zinc oxide 223–229 nanorod surface-enhanced Raman scattering substrates without [122] Menazea A A, Abdelghany A M, Hakeem N A, et al. and with gold nanoparticles fabricated through pulsed-laser- Precipitation of silver nanoparticles in borate glasses by induced photolysis. Applied Sciences, 2020, 10(14): 5015 1064 nm Nd:YAG nanosecond laser pulses: characterization [108] Yahata A, Ishii H, Nakamura K, et al. Three-dimensional and dielectric studies. Journal of Electronic Materials, 2020, periodic structures of gold nanoclusters in the interstices of sub- 49(1): 826–832 100 nm polymer particles toward surface-enhanced Raman [123] Menazea A A, Abdelghany A M, Hakeem N A, et al. Nd:YAG scattering. Advanced Powder Technology, 2019, 30(12): nanosecond laser pulses for precipitation silver nanoparticles in 2957–2963 silicate glasses: AC conductivity and dielectric studies. Silicon, [109] Belusso L C S, Lenz G F, Fiorini E E, et al. Synthesis of silver 2020, 12(1): 13–20 nanoparticles from bottom up approach on borophosphate glass [124] Abdel-Khalek E K, Elsharkawy M A, Motawea M A, et al. and their applications as SERS, antibacterial and glass-based Dielectric and thermal properties of tetragonal PbTiO3 catalyst. Applied Surface Science, 2019, 473: 303–312 nanoparticles/clusters embedded in lithium tetraborate glass [110] Goodenough J B, Braga M H. Batteries for electric road [111] Braga M H, Grundish N S, Murchison A J, et al. Alternative strategy for a safe rechargeable battery. Energy matrix. Silicon, 2021, 13: 2993–3002 [125] Rejikumar P R, Jyothy P V, Mathew S, et al. Effect of silver vehicles. Dalton Transactions, 2018, 47(3): 645–648 nanoparticles on the dielectric properties of holmium doped & silica glass. Physical Review B: Condensed Matter, 2010, 405: [112] Mikolajczak C, Kahn M, White K, et al. Lithium-Ion Batteries [126] Abdelghany A M, Zeyada H M, ElBatal H A, et al. AC Hazard and Use Assessment. Springer, 2011, ISBN-13: conductivity and dielectric behavior of silicophosphate glass 978–1461434856 doped by Nd2O3. Silicon, 2017, 9(3): 347–354 Environmental Science, 2017, 10(1): 331–336 1513–1517 [113] Braga M H, Murchison A J, Ferreira J A, et al. Glass- [127] Yang K, Liu J, Shen B, et al. Large improvement on energy amorphous alkali-ion solid electrolytes and their performance storage and charge‒discharge properties of Gd2O3-doped in symmetrical cells. Energy & Environmental Science, 2016, BaO‒K2O‒Nb2O5‒SiO2 glass-ceramic dielectrics. Materials 9(3): 948–954 Science and Engineering B, 2017, 223: 178–184 [114] Kamaya N, Homma K, Yamakawa Y, et al. A lithium superionic conductor. Nature Materials, 2011, 10(9): 682–686 [115] Duclot M, Souquet J L. Glassy materials for lithium batteries: electrochemical properties and devices performances. Journal of Power Sources, 2001, 97‒98: 610–615 [116] Janek J, Zeier W G. A solid future for battery development. Nature Energy, 2016, 1(9): 16141 [117] Manuel Stephan A, Nahm K S. Review on composite polymer [128] Fonseca J, Choi S. Electro-and photoelectro-catalysts derived from bimetallic amorphous metal-organic frameworks. Catalysis Science & Technology, 2020, 10(24): 8265–8282 [129] Ellis L D, Rorrer N A, Sullivan K P, et al. Chemical and biological catalysis for plastics recycling and upcycling. Nature Catalysis, 2021, 4(7): 539–556 [130] Schlögl R. Heterogeneous catalysis. Angewandte Chemie International Edition, 2015, 54(11): 3465–3520 Javier FONSECA. NPs embedded into glass matrices: GNCs 71 [131] Gärtner D, Sandl S, Jacobi von Wangelin A. Homogeneous vs. [145] Baino F, Novajra G, Miguez-Pacheco V, et al. Bioactive heterogeneous: mechanistic insights into iron group metal- glasses: special applications outside the skeletal system. Journal catalyzed reductions from poisoning experiments. Catalysis of Non-Crystalline Solids, 2016, 432: 15–30 Science & Technology, 2020, 10(11): 3502–3514 [146] Hench L L, Splinter R J, Allen W C, et al. Bonding [132] Lee A F, Bennett J A, Manayil J C, et al. Heterogeneous mechanisms at the interface of ceramic prosthetic materials. catalysis for sustainable biodiesel production via esterification Journal of Biomedical Materials Research, 1971, 5(6): 117–141 and transesterification. Chemical Society Reviews, 2014, [147] Erol-Taygun M, Zheng K, Boccaccini A R. Nanoscale 43(22): 7887–7916 bioactive glasses in medical applications. International Journal [133] Khan S A, Khan N, Irum U, et al. Cellulose acetate-Ce/Zr@Cu0 catalyst for the degradation of organic pollutant. International Journal of Biological Macromolecules, 2020, 153: 806–816 [134] Gao C, Lyu F, Yin Y. Encapsulated metal nanoparticles for catalysis. Chemical Reviews, 2021, 121(2): 834–881 nanoclusters catalysis with by atomically crystallographic precise structures. synthesis to materials design for biomedical applications. Materials, 2016, 9(4): 288–295 [149] Hum J, Boccaccini A R. Bioactive glasses as carriers for [135] Jin R, Li G, Sharma S, et al. Toward active-site tailoring in heterogeneous of Applied Glass Science, 2013, 4(2): 136–148 [148] Vichery C, Nedelec J M. Bioactive glass nanoparticles: from metal Chemical Reviews, 2021, 121(2): 567–648 bioactive molecules and therapeutic drugs: a review. Journal of Materials Science: Materials in Medicine, 2012, 23(10): 2317–2333 [150] Wu C, Fan W, Chang J. Functional mesoporous bioactive glass [136] Majhi S, Sharma K, Singh R, et al. Development of silver nanospheres: synthesis, high loading efficiency, controllable nanoparticles decorated on functional glass slide as highly delivery of doxorubicin and inhibitory effect on bone cancer efficient and recyclable dip catalyst. ChemistrySelect, 2020, cells. Journal of Materials Chemistry B, 2013, 1(21): 5(40): 12365–12370 2710–2718 [137] Mennecke K, Cecilia R, Glasnov T N, et al. Palladium(0) [151] Blackburn G, Scott T G, Bayer I S, et al. Bionanomaterials for nanoparticles on glass‒polymer composite materials as bone tumor engineering and tumor destruction. Journal of recyclable catalysts: a comparison study on their use in batch Materials Chemistry B, 2013, 1(11): 1519–1534 and continuous flow processes. Advanced Synthesis & Catalysis, 2008, 350(5): 717–730 [138] Elhage A, Wang B, Marina N, et al. Glass wool: a novel support for heterogeneous catalysis. Chemical Science, 2018, 9(33): 6844–6852 [139] Lenz G F, Schneider R, Rossi de Aguiar K M F, et al. Selfsupported nickel nanoparticles on germanophosphate glasses: synthesis and applications in catalysis. RSC Advances, 2019, 9(30): 17157–17164 [152] Jayalekshmi A C, Victor S P, Sharma C P. Magnetic and degradable polymer/bioactive glass composite nanoparticles for biomedical applications. Colloids and Surfaces B: Biointerfaces, 2013, 101: 196–204 [153] Wu C, Fan W, Zhu Y, et al. Multifunctional magnetic mesoporous bioactive glass scaffolds with a hierarchical pore structure. Acta Biomaterialia, 2011, 7(10): 3563–3572 [154] Shih S J, Tzeng W L, Jatnika R, et al. Control of Ag nanoparticle distribution influencing bioactive and antibacterial [140] Luo X, Meng M, Li R, et al. Honeycomb-like porous metallic properties of Ag-doped mesoporous bioactive glass particles glasses decorated by Cu nanoparticles formed by one-pot prepared by spray pyrolysis. Journal of Biomedical Materials electrochemically galvanostatic etching. Materials & Design, Research Part B: Applied Biomaterials, 2015, 103(4): 899–907 2020, 196: 109109 [141] Zheng K, Boccaccini A R. Sol-gel processing of bioactive glass [155] Li J, Zhai D, Lv F, et al. Preparation of copper-containing bioactive glass/eggshell membrane nanocomposites for nanoparticles: a review. Advances in Colloid and Interface improving angiogenesis, antibacterial activity and wound Science, 2017, 249: 363–373 healing. Acta Biomaterialia, 2016, 36: 254–266 [142] Brauer D S. Bioactive glasses — structure and properties. [156] Ventura M G, Parola A J, Pires de Matos A. Influence of heat Angewandte Chemie International Edition, 2015, 54(14): treatment on the colour of Au and Ag glasses produced by the 4160–4181 sol-gel pathway. Journal of Non-Crystalline Solids, 2011, [143] Jones J R. Review of bioactive glass: from Hench to hybrids. Acta Biomaterialia, 2013, 9(1): 4457–4486 [144] Miguez-Pacheco V, Hench L L, Boccaccini A R. Bioactive glasses beyond bone and teeth: emerging applications in contact with soft tissues. Acta Biomaterialia, 2015, 13: 1–15 357(4): 1342–1349 [157] Mazzold P, Carturan S, Quaranta A, et al. Ion exchange process: history, evolution and applications. Rivista del Nuovo Cimento, 2013, 36: 397–460 [158] Molina G, Murcia S, Molera J, et al. Color and dichroism of 72 Front. Mater. Sci. 2022, 16(3): 220607 silver-stained glasses. Journal of Nanoparticle Research, 2013, [172] Singh S P, Karmakar B. Single-step synthesis and surface plasmons of bismuth-coated spherical to hexagonal silver 15(9): 1932 [159] Pradell T, Molera J, Roque J, et al. Ionic-exchange mechanism in the formation of medieval luster decorations. Journal of the nanoparticles in dichroic Ag:bismuth glass nanocomposites. Plasmonics, 2011, 6(3): 457–467 [173] Venkateswara Rao G, Shashikala H D. Effect of heat treatment American Ceramic Society, 2005, 88(5): 1281–1289 [160] Palomar T, Redol P, Cruz Almeida I, et al. The influence of on optical, dielectric and mechanical properties of silver environment in the alteration of the stained-glass windows in nanoparticle embedded CaO‒CaF2‒P2O5 glass. Journal of Portuguese monuments. Heritage, 2018, 1(2): 365–376 Alloys and Compounds, 2015, 622: 108–114 [161] Palomar T, Grazia C, Pombo Cardoso I, et al. Analysis of [174] Swetha B N, Keshavamurthy K. Impact of thermal annealing visible time on luminescence properties of Eu3+ ions in silver hyperspectral imaging in-situ. Spectrochimica Acta Part A: nanoparticles embedded lanthanum sodium borate glasses. Molecular and Biomolecular Spectroscopy, 2019, 223: 117378 Applied Surface Science, 2020, 525: 146505 chromophores in stained-glass windows using [162] Machado C, Machado A, Palomar T, et al. Debitus grisailles for [175] Mohammed Danmallam I, Ghoshal S K, Ariffin R, et al. stained-glass conservation: an analytical study. Conservar Europium luminescence in silver and gold nanoparticles co- Património, 2020, 34: 65–72 embedded phosphate glasses: Judd‒Ofelt calculation. Optical [163] Gonella F, Mazzoldi P. Chapter 2: Metal nanocluster composite glasses. Handbook of Nanostructured Materials and Materials, 2020, 105: 109889 [176] Qiu J, Shirai M, Nakaya T, et al. Space-selective precipitation of metal nanoparticles inside glasses. Applied Physics Letters, Nanotechnology, 2000, 4: 81–158 [164] Liu L C, Risbud S H. Quantum-dot size-distribution analysis and precipitation stages in semiconductor doped glasses. 2002, 81(16): 3040–3042 [177] Qiu J, Jiang X, Zhu C, et al. Manipulation of gold nanoparticles inside transparent materials. Angewandte Chemie International Journal of Applied Physics, 1990, 68(1): 28–32 [165] Machado T M, Falci R F, Silva I L, et al. Erbium 1.55 µm Edition, 2004, 43(17): 2230–2234 luminescence enhancement due to copper nanoparticles [178] Hu X, Zhao Q, Jiang X, et al. Space-selective co-precipitation plasmonic activity in tellurite glasses. Materials Chemistry and of silver and gold nanoparticles in femtosecond laser pulses Physics, 2019, 224: 73–78 irradiated Ag+, Au3+ co-doped silicate glass. Solid State [166] Machado T M, Falci R F, Andrade G F S, et al. Unprecedented Communications, 2006, 138(1): 43–46 multiphonon vibronic transitions of erbium ions on copper [179] Teng Y, Qian B, Jiang N, et al. Light and heat driven nanoparticle-containing tellurite glasses. Physical Chemistry precipitation of copper nanoparticles inside Cu2+-doped borate glasses. Chemical Physics Letters, 2010, 485(1‒3): 91–94 Chemical Physics, 2020, 22(23): 13118–13122 [167] Rajaramakrishna R, Saiyasombat C, Anavekar R V, et al. [180] Almeida J M P, De Boni L, Avansi W, et al. Generation of Structure and nonlinear optical studies of Au nanoparticles copper nanoparticles induced by fs-laser irradiation in embedded in lead lanthanum borate glass. Journal of Non- borosilicate glass. Optics Express, 2012, 20(14): 15106–15113 [181] Chen S, Akai T, Kadono K, et al. Reversible control of silver Crystalline Solids, 2014, 406: 107–110 [168] Rajaramakrishna R, Karuthedath S, Anavekar R V, et al. nanoparticle generation and dissolution in soda-lime silicate Nonlinear optical studies of lead lanthanum borate glass doped glass through X-ray irradiation and heat treatment. Applied with Au nanoparticles. Journal of Non-Crystalline Solids, 2012, Physics Letters, 2001, 79(22): 3687–3689 [182] Sheng J, Kadono K, Yazawa T. Nanosized gold clusters 358(14): 1667–1672 [169] Chen F, Dai S, Xu T, et al. Surface-plasmon enhanced ultrafast third-order optical nonlinearities in ellipsoidal gold nanoparticles embedded bismuthate glasses. Chemical Physics nanoclusters in oxyfluoride glass. Journal of Luminescence, 2014, 152: 222–225 [171] Guo Z, Ye S, Liu T, et al. SmF3 doping and heat treatment manipulated Ag species evaluation and efficient energy transfer from Ag nanoclusters to Sm3+ of Non-Crystalline Solids, 2003, 324(3): 295–299 [183] Valentin E, Bernas H, Ricolleau C, et al. Ion beam “photography”: decoupling nucleation and growth of metal Letters, 2011, 514(1‒3): 79–82 [170] Ma R, Qian J, Cui S, et al. Enhancing NIR emission of Yb3+ by silver formation in selected areas of soda-lime silicate glass. Journal ions in oxyfluoride glass. Journal of Non-Crystalline Solids, 2017, 458: 80–85 clusters in glass. Physical Review Letters, 2001, 86(1): 99–102 [184] Farag H K, Marzouk M A. Preparation and characterization of nanostructured nickel oxide and its influence on the optical properties of sodium zinc borate glasses. Journal of Materials Science: Materials in Electronics, 2017, 28(20): 15480–15487 [185] Asyikin A S, Halimah M K, Latif A A, et al. Physical, structural and optical properties of bio-silica borotellurite glass Javier FONSECA. NPs embedded into glass matrices: GNCs system doped with samarium oxide nanoparticles. Journal of 73 of multicore magnetite nanoparticles crystallised from a silicate glass. Journal of Materials Science, 2012, 47(15): 5886–5890 Non-Crystalline Solids, 2020, 529: 119777 [186] Orives J R, Viali W R, Magnani M, et al. Incorporation of [199] Woltz S, Hiergeist R, Görnert P, et al. Magnetite nanoparticles CdFe2O4‒SiO2 nanoparticles in SbPO4‒ZnO‒PbO glasses by prepared by the glass crystallization method and their physical melt-quenching process. Eclética Química Journal, 2018, 43(2): properties. Journal of Magnetism and Magnetic Materials, 2006, 298(1): 7–13 32–43 [187] Widanarto W, Sahar M R, Ghoshal S K, et al. Thermal, [200] Burda C, Chen X, Narayanan R, et al. Chemistry and properties structural and magnetic properties of zinc-tellurite glasses of nanocrystals of different shapes. Chemical Reviews, 2005, containing natural ferrite oxide. Materials Letters, 2013, 108: 105(4): 1025–1102 [201] Maurer R D. Nucleation and growth in a photosensitive glass. 289–292 [188] Widanarto W, Sahar M R, Ghoshal S K, et al. Effect of natural Journal of Applied Physics, 1958, 29(1): 1–8 Fe3O4 nanoparticles on structural and optical properties of Er3+ [202] Zhang K, Zhou D, Qiu J, et al. Silver nanoparticles enhanced doped tellurite glass. Journal of Magnetism and Magnetic luminescence and stability of CsPbBr3 perovskite quantum dots Materials, 2013, 326: 123–128 in borosilicate glass. Journal of the American Ceramic Society, [189] Anigrahawati P, Sahar M R, Ghoshal S K. Influence of Fe3O4 2020, 103(4): 2463–2470 nanoparticles on structural, optical and magnetic properties of [203] Manzani D, Almeida J M P, Napoli M, et al. Nonlinear optical erbium doped zinc phosphate glass. Materials Chemistry and properties of tungsten lead-pyrophosphate glasses containing Physics, 2015, 155: 155–161 metallic [190] Franco D F, Sant’Ana A C, De Oliveira L F C, et al. The Sb2O3 copper nanoparticles. Plasmonics, 2013, 8(4): 1667–1674 redox route to obtain copper nanoparticles in glasses with [204] Qiu J, Zhu C, Nakaya T, et al. Space-selective valence state plasmonic properties. Journal of Materials Chemistry C, 2015, manipulation of transition metal ions inside glasses by a 3(15): 3803–3808 femtosecond laser. Applied Physics Letters, 2001, 79(22): [191] Egorova O N, Semjonov S L, Velmiskin V V, et al. Phosphatecore silica-clad Er/Yb-doped optical fiber and cladding pumped 3567–3569 [205] Zeng H, Qiu J, Ye Z, et al. Irradiation assisted fabrication of gold nanoparticles-doped glasses. Journal of Crystal Growth, laser. Optics Express, 2014, 22(7): 7632–7637 [192] Sathi Z M, Zhang J, Luo Y, et al. Improving broadband emission within Bi/Er doped silicate fibres with Yb co-doping. 2004, 267(1‒2): 156–160 [206] Hosono H, Matsunami N, Kudo A, et al. Novel approach for synthesizing Ge fine particles embedded in glass by ion Optical Materials Express, 2015, 5(10): 2096–2105 [193] Wei T, Chen F, Tian Y, et al. Broadband near-infrared emission implantation: formation of Ge nanocrystal in SiO2‒GeO2 property in Er3+/Ce3+ co-doped silica-germanate glass for fiber glasses by proton implantation. Applied Physics Letters, 1994, amplifier. Spectrochimica Acta Part A: Molecular and 65(13): 1632–1634 [207] Hosono H, Kawamura K, Kameshima Y, et al. Nanometer- Biomolecular Spectroscopy, 2014, 126: 53–58 [194] Chen D D, Qian Q, Peng M Y, et al. Effect of Ce3+, Dy3+, and sized Ge particles in GeO2–SiO2 glasses produced by proton Tb3+ additions on the spectroscopic properties of Er3+/Yb3+ implantation: combined effects of electronic excitation and codoped tellurite glasses. Physica B: Condensed Matter, 2010, chemical reaction. Journal of Applied Physics, 1997, 82(9): 405(21): 4453–4456 4232–4235 Eu2+ on the silver [208] Lifshitz I M, Slyozov V V. The kinetics of precipitation from aggregates formation in Ag+–Na+ ion-exchanged Eu3+-doped supersaturated solid solutions. Journal of Physics and [195] Li L, Yang Y, Zhou D, et al. Influence of the sodium-aluminosilicate glasses. Journal of the American [196] Worsch C, Büttner M, Schaaf P, et al. Magnetic properties of multicore magnetite nanoparticles prepared Chemistry of Solids, 1961, 19(1‒2): 35–50 [209] Houk L R, Challa S R, Grayson B, et al. The definition of Ceramic Society, 2014, 97(4): 1110–1114 by glass crystallization. Journal of Materials Science, 2013, 48(6): 2299–2307 [197] Harizanova R, Völksch G, Rüssel C. Microstructures formed during devitrification of Na2O·Al2O3·B2O3·SiO2·Fe2O3 glasses. Journal of Materials Science, 2010, 45(5): 1350–1353 [198] Worsch C, Schaaf P, Harizanova R, et al. Magnetisation effects “critical radius” for a collection of nanoparticles undergoing Ostwald ripening. Langmuir, 2009, 25(19): 11225–11227 [210] Azlina Y, Azlan M N, Halimah M K, et al. Optical performance of neodymium nanoparticles doped tellurite glasses. Physica B: Condensed Matter, 2020, 577: 411784 [211] Muhammad Noorazlan A, Mohamed Kamari H, Zulkefly S S, et al. Effect of erbium nanoparticles on optical properties of zinc borotellurite glass system. Journal of Nanomaterials, 2013, 74 Front. Mater. Sci. 2022, 16(3): 220607 2013: 940917 Science and Technology, 1996, 6(3): 203–217 [212] Halimah M K, Awshah A A, Hamza A M, et al. Effect of [227] Kajihara K. Recent advances in sol-gel synthesis of monolithic neodymium nanoparticles on optical properties of zinc tellurite silica and silica-based glasses. Journal of Asian Ceramic glass system. Journal of Materials Science: Materials in Electronics, 2020, 31(5): 3785–3794 [213] Orives J R, Viali W R, Santagneli S H, et al. Phosphate glasses Societies, 2013, 1(2): 121–133 [228] Adachi T, Sakka S. Preparation of monolithic silica gel and glass by the sol-gel method using N, N-dimethylformamide. via coacervation route containing CdFe2O4 nanoparticles: Journal of Materials Science, 1987, 22(12): 4407–4410 structural, optical and magnetic characterization. Dalton [229] Chan J B, Jonas J. Effect of various amide additives on the transactions, 2018, 47(16): 5771–5779 [214] Orives J R, Pichon B P, Mertz D, et al. Phosphate glasses tetramethoxysilane sol-gel process. Journal of Non-Crystalline Solids, 1990, 126(1‒2): 79–86 containing monodisperse Fe3−δO4@SiO2 stellate nanoparticles [230] Kirkbir F, Murata H, Meyers D, et al. Drying of large obtained by melt-quenching process. Ceramics International, monolithic aerogels between atmospheric and supercritical 2020, 46(8): 12120–12127 pressures. Journal of Sol-Gel Science and Technology, 1998, [215] Hench L L, West J K. The sol-gel process. Chemical Reviews, 1990, 90(1): 33–72 13(1/3): 311–316 [231] Iler R K. The Chemistry of Silica: Solubility, Polymerization, [216] Graham T. On the properties of silicic acid and other analogous colloidal substances. Journal of the Chemical Society, 1864, 17: 318–327 Colloid and Surface Properties, and Biochemistry. New York, USA: John Wiley and Sons Ltd., 1979 [232] Yamane M, Aso S, Okano S, et al. Low temperature synthesis [217] Dislich H. New routes to multicomponent oxide glasses. Angewandte Chemie International Edition, 1971, 10(6): 363–370 of a monolithic silica glass by the pyrolysis of a silica gel. Journal of Materials Science, 1979, 14(3): 607–611 [233] Hæreid S, Einarsrud M A, Scherer G W. Mechanical [218] Yoldas B E. Monolithic glass formation by chemical strengthening of TMOS-based alcogels by aging in silane polymerization. Journal of Materials Science, 1979, 14(8): solutions. Journal of Sol-Gel Science and Technology, 1994, 3: 1843–1849 199–204 [219] Razdobreev I, El Hamzaoui H, Ivanov V Y, et al. Optical [234] Kawaguchi T, Hishikura H, Iura J, et al. Monolithic dried gels spectroscopy of bismuth-doped pure silica fiber preform. Optics and silica glass prepared by the sol-gel process. Journal of Non- Letters, 2010, 35(9): 1341–1343 [220] Vallet-Regí M. Ceramics for medical applications. Journal of the Chemical Society — Dalton Transactions: Inorganic Chemistry, 2001, 2(2): 97–108 Crystalline Solids, 1984, 63(1‒2): 61–69 [235] Sarkar A, Chaudhuri S R, Wang S, et al. Drying of alkoxide gels-observation of an alternate phenomenology. Journal of Sol-Gel Science and Technology, 1994, 2(1‒3): 865–870 [221] Kaur G, Pickrell G, Sriranganathan N, et al. Review and the [236] Murata H, Meyers D E, Kirkbir F, et al. Drying and sintering of state of the art: sol-gel and melt quenched bioactive glasses for bulk silica gels. Journal of Sol-Gel Science and Technology, tissue engineering. Journal of Biomedical Materials Research Part B: Applied Biomaterials, 2016, 104(6): 1248–1275 [222] El Hamzaoui H, Bigot L, Bouwmans G, et al. From molecular 1997, 8(1‒3): 397–402 [237] Scherer G W, Smith D M. Cavitation during drying of a gel. Journal of Non-Crystalline Solids, 1995, 189(3): 197–211 precursors in solution to microstructured optical fiber: a sol-gel [238] Smith D M, Stein D, Anderson J M, et al. Preparation of low- polymeric route. Optical Materials Express, 2011, 1(2): density xerogels at ambient pressure. Journal of Non- 234–242 Crystalline Solids, 1995, 186: 104–112 [223] Brinker C J, Scherer G W. Sol-Gel Science: The Physics and [239] Susa K, Matsuyama I, Satoh S, et al. Sol-gel derived Ge-doped Chemistry of Sol-Gel Processing. New York, USA: Academic silica glass for optical fiber application: I. Preparation of gel Press, 1990 and glass and their characterization. Journal of Non-Crystalline [224] Zarzycki J, Prassas M, Phalippou J. Synthesis of glasses from gels: the problem of monolithic gels. Journal of Materials Science, 1982, 17(11): 3371–3379 [225] Ulrich D R. Prospects of sol-gel processes. Journal of NonCrystalline Solids, 1988, 100(1‒3): 174–193 Solids, 1990, 119(1): 21–28 [240] Susa K, Matsuyama I, Satoh S. Sol-gel derived Ge-doped silica glass for optical fiber application. II. Excess optical loss. Journal of Non-Crystalline Solids, 1991, 128(2): 118–125 [241] Breitscheidel B, Zieder J, Schubert U. Metal complexes in [226] Kirkbir F, Murata H, Meyers D, et al. Drying and sintering of inorganic matrices. 7. Nanometer-sized, uniform metal particles sol-gel derived large SiO2 monoliths. Journal of Sol-Gel in a silica matrix by sol-gel processing of metal complexes. Javier FONSECA. NPs embedded into glass matrices: GNCs Chemistry of Materials, 1991, 3(3): 559–566 [242] Lembacher C, Schubert U. Nanosized platinum particles by sol- 75 behavior-enhanced optical nonlinearities. Journal of NonCrystalline Solids, 2017, 459: 142–149 gel processing of tethered metal complexes: influence of the [255] Zhang Y, Jin Y, He M, et al. Optical properties of bimetallic precursors and the organic group removal method on the Au‒Cu nanocrystals embedded in glass. Materials Research particle size. New Journal of Chemistry, 1998, 22(7): 721–724 Bulletin, 2018, 98: 94–102 [243] Trimmel G, Lembacher C, Kickelbick G, et al. Sol-gel [256] Le Rouge A, El Hamzaoui H, Capoen B, et al. Synthesis and processing of alkoxysilyl-substituted nickel complexes for the nonlinear preparation of highly dispersed nickel in silica. New Journal of nanoparticles embedded in sol-gel derived silica glass. Chemistry, 2002, 26(6): 759–765 optical properties of zirconia-protected gold Materials Research Express, 2015, 2(5): 055009 [244] Lukehart C M, Milne S B, Stock S R. Formation of crystalline [257] Campos-Zuñiga E E, Alonso-Lemus I L, Agarwal V, et al. Sol- nanoclusters of Fe2P, RuP, Co2P, Rh2P, Ni2P, Pd5P2, or PtP2 in gel synthesis for stable green emission in samarium doped a silica xerogel matrix from single-source molecular precursors. borosilicate glasses. Ceramics International, 2019, 45(18): Chemistry of Materials, 1998, 10(3): 903–908 24052–24059 [245] Carpenter J P, Lukehart C M, Stock S R, et al. Formation of a [258] Schubert U, Amberg-Schwab S, Breitscheidel B. Metal nanocomposite containing particles of Co3C from a single- complexes in inorganic matrices. 4. Small metal particles in source precursor bound to a silica xerogel host matrix. palladium‒silica composites by sol-gel processing of metal Chemistry of Materials, 1995, 7(1): 201–205 [246] Carpenter J P, Lukehart C M, Henderson D O, et al. Formation complexes. Chemistry of Materials, 1989, 1(6): 576–578 [259] Schubert U. Metal oxide/silica and metal/silica nanocomposites of crystalline germanium nanoclusters in a silica xerogel matrix from from an organogermanium precursor. Chemistry of Materials, Advanced Engineering Materials, 2004, 6(3): 173–176 1996, 8(6): 1268–1274 [247] Carpenter J P, Lukehart C M, Milne S B, et al. Organometallic compounds as single-source precursors to nanocomposite materials: an overview. Journal of Organometallic Chemistry, 1998, 557(1): 121–130 [248] Mathieu H, Richard T, Allègre J, et al. Quantum confinement effects of CdS nanocrystals in a sodium borosilicate glass prepared by the sol-gel process. Journal of Applied Physics, 1995, 77(1): 287–293 organofunctional single-source sol-gel precursors. [260] Schubert U. Preparation of metal oxide or metal nanoparticles in silica via metal coordination to organofunctional trialkoxysilanes. Polymer International, 2009, 58(3): 317–322 [261] Trimmel G, Schubert U. Sol-gel processing of tethered metal complexes: influence of the metal and the complexing alkoxysilane on the texture of the obtained silica gels. Journal of Non-Crystalline Solids, 2001, 296(3): 188–200 [262] Moerke W, Lamber R, Schubert U, et al. Metal complexes in inorganic matrices. 11. Composition of highly dispersed [249] Zhong J, Ma X, Lu H, et al. Preparation and optical properties bimetallic Ni, Pd alloy particles prepared by sol-gel processing: of sodium borosilicate glasses containing Sb nanoparticles. electron microscopy and FMR study. Chemistry of Materials, Journal of Alloys and Compounds, 2014, 607: 177–182 1994, 6(10): 1659–1666 [250] Zhong J, Xiang W, Zhao H, et al. Synthesis, characterization, [263] Kaiser A, Görsmann C, Schubert U. Influence of the metal and third-order nonlinear optical properties of copper quantum complexation on size and composition of Cu/Ni nano-particles dots embedded in sodium borosilicate glass. Journal of Alloys prepared by sol-gel processing. Journal of Sol-Gel Science and and Compounds, 2012, 537: 269–274 Technology, 1997, 8(1‒3): 795–799 [251] Pei L, Liang X, Cai W, et al. Photocatalytic activity in [264] Malenovska M, Neouze M A, Schubert U, et al. Multi- monodisperse In2O3 nanocrystals incorporated into transparent component hybrid organic‒inorganic particles with highly silica glassy matrix. Journal of Alloys and Compounds, 2015, dispersed silver nanoparticles in the external shell. Dalton 626: 131–135 Transactions, 2008(34): 4647–4651 [252] Gao H, Xiang W, Ma X, et al. Sol-gel synthesis and third-order optical nonlinearity of Au nanoparticles doped monolithic glass. Gold Bulletin, 2015, 48(3‒4): 153–159 [265] Reisfeld R. Prospects of sol-gel technology towards luminescent materials. Optical Materials, 2001, 16(1‒2): 223707 [253] Zhong J, Xiang W. Sol-gel synthesis and third-order nonlinear [266] Yang G, Cheng S, Li C, et al. Investigation of the oxidation optical properties of Cu3.8Ni nanoparticles doped glass. Journal states of Cu additive in colored borosilicate glasses by electron of Non-Crystalline Solids, 2017, 462: 17–22 energy loss spectroscopy. Journal of Applied Physics, 2014, [254] Huang Y, Xiang W, Lin S, et al. The synthesis of bimetallic gold plus nickel nanoparticles dispersed in a glass host and 116(22): 1–6 [267] Yanes A C, Santana-Alonso A, Méndez-Ramos J, et al. Novel 76 Front. Mater. Sci. 2022, 16(3): 220607 sol-gel nano-glass-ceramics comprising Ln3+-doped YF3 nanocrystals: structure and high efficient UV up-conversion. [268] Zhang M, Fan H, Xi B, et al. Synthesis, characterization, and properties of uniform Ln3+-doped 115–130 [282] De Marchi G, Mattei G, Mazzoldi P, et al. Two stages in the Advanced Functional Materials, 2011, 21(16): 3136–3142 luminescence Nanocomposites Synthesis, Properties and Applications, 2016: YF3 nanospindles. The Journal of Physical Chemistry C, 2007, kinetics of gold cluster growth in ion-implanted silica during isothermal annealing in oxidizing atmosphere. Journal of Applied Physics, 2002, 92(8): 4249–4254 [283] Srivastava S K, Gangopadhyay P, Amirthapandian S, et al. 111(18): 6652–6657 [269] Wang G, Qin W, Zhang J, et al. Synthesis, growth mechanism, Effects of high-energy Si ion-irradiations on optical responses Yb3+/Tm3+-codoped of Ag metal nanoparticles in a SiO2 matrix. Chemical Physics and tunable upconversion luminescence of YF3 nanobundles. The Journal of Physical Chemistry C, 2008, Letters, 2014, 607: 100–104 [284] Liu Q, He X, Zhou X, et al. Third-order nonlinearity in Ag- 112(32): 12161–12167 [270] Kalwarczyk E, Kabaciński P, Kardaś T M, et al. A Seedless nanoparticles embedded 56GeS2–24Ga2S3–20KBr chalcohalide method for gold nanoparticle growth inside a silica matrix: glasses. Journal of Non-Crystalline Solids, 2011, 357(11‒13): fabrication of materials capable of third-harmonic generation in 2320–2323 [285] Vytykacova S, Svecova B, Nekvindova P, et al. The formation the near-infrared. ChemPlusChem, 2019, 84(5): 525–533 [271] Melnikov P, Arkhangelsky I V, Nascimento V A, et al. of silver metal nanoparticles by ion implantation in silicate Thermolysis mechanism of samarium nitrate hexahydrate. glasses. Nuclear Instruments & Methods in Physics Research Journal of Thermal Analysis and Calorimetry, 2014, 118(3): Section B: Beam Interactions with Materials and Atoms, 2016, 371: 245–250 1537–1541 [272] Gaddam A, Fernandes H R, Tulyaganov D U, et al. The [286] Zhang B, Sato R, Oyoshi K, et al. Dispersion of third-order structural role of lanthanum oxide in silicate glasses. Journal of susceptibility Non-Crystalline Solids, 2019, 505: 18–27 implantation. Nuclear Instruments & Methods in Physics [273] Harper C A. Handbook of Ceramics, Glasses, and Diamonds.New York, USA: McGraw-Hill, 2001 [274] Pope E J A, Mackenzie J D. Nd-doped silica glass I: structural of Au nanoparticles fabricated by ion Research Section B: Beam Interactions with Materials and Atoms, 2019, 447: 38–42 [287] Herrera A, Balzaretti N M. Effect of gold nanoparticles in broadband 1988, 106(1‒3): 236–241 B2O3‒PbO‒Bi2O3‒GeO2 glass. Journal of Luminescence, 2017, [275] Campostrini R, Carturan G, Ferrari M, et al. Luminescence of Eu3+ ions during thermal densification of SiO2 gel. Journal of Materials Research, 1992, 7(3): 745–753 near-infrared emission of Pr3+ evolution in the sol-gel state. Journal of Non-Crystalline Solids, doped 181: 147–152 [288] Stepanov A L, Galyautdinov M F, Evlyukhin A B, et al. Synthesis of periodic plasmonic microstructures with copper [276] Pope E J A, Mackenzie J D. Sol-gel processing of neodymian- nanoparticles in silica glass by low-energy ion implantation. silica glass. Journal of the American Ceramic Society, 1993, Applied Physics A: Materials Science & Processing, 2013, 76(5): 1325–1328 111(1): 261–264 [277] Chakrabarti S, Sahu J, Chakraborty M, et al. Monophasic silica [289] Ghosh B, Chakraborty P, Singh B P, et al. Enhanced nonlinear glasses with large neodymia concentration. Journal of Non- optical responses in metal–glass nanocomposites. Applied Crystalline Solids, 1994, 180(1): 96–101 Surface Science, 2009, 256(2): 389–394 [278] Monteil A, Chaussedent S, Alombert-Goget G, et al. Clustering [290] Wang Y H, Wang Y M, Lu J D, et al. Nonlinear optical of rare earth in glasses, aluminum effect: experiments and properties of Cu nanoclusters by ion implantation in silicate modeling. Journal of Non-Crystalline Solids, 2004, 348: 44–50 glass. Optics Communications, 2010, 283(3): 486–489 [279] Langlet M, Coutier C, Meffre W, et al. Microstructural and [291] Xiao X H, Ren F, Wang J B, et al. Formation of aligned silver spectroscopic study of sol-gel derived Nd-doped silica glasses. nanoparticles by ion implantation. Materials Letters, 2007, Journal of Luminescence, 2002, 96(2‒4): 295–309 61(22): 4435–4437 [280] Sckerl M W, Guldberg-Kjaer S, Rysholt Poulsen M, et al. [292] Shen Y, Qi T, Qiao Y, et al. The effect of Ni pre-implantation Precipitate coarsening and self organization in erbium-doped on surface morphology and optical absorption properties of Ag silica. Physical Review B: Condensed Matter, 1999, 59(21): nanoparticles embedded in SiO2. Applied Surface Science, 13494–13497 2016, 363: 310–317 [281] Stepanov A L. Chapter 4: Laser annealing of metal [293] Wang J, Jia G, Zhang B, et al. Formation and optical absorption nanoparticles synthesized in glasses by ion implantation. Glass property of nanometer metallic colloids in Zn and Ag dually Javier FONSECA. NPs embedded into glass matrices: GNCs implanted silica: synthesis of the modified Ag nanoparticles. 77 implantation. Journal of Materials Science and Technology, 2009, 25: 669–672 Journal of Applied Physics, 2013, 113(3): 034304 [294] Nistor L C, van Landuyt J, Barton J B, et al. Colloid size distribution in ion implanted glass. Journal of Non-Crystalline [307] Bourgoin J C, Corbett J W. Enhanced diffusion mechanisms. Radiation Effects, 1978, 36(3‒4): 157–188 [308] Jia G, Xu R, Mu X, et al. Zn ion post-implantation-driven Solids, 1993, 162(3): 217–224 [295] Stepanov A L, Popok V N. Nanostructuring of silicate glass synthesis of CuZn alloy nanoparticles in Cu-preimplanted silica under low-energy Ag-ion implantation. Surface Science, 2004, and their thermal evolution. ACS Applied Materials & 566‒568: 1250–1254 Interfaces, 2013, 5(24): 13055–13062 [296] Gnaser H, Brodyanski A, Reuscher B. Focused ion beam [309] Wang J, Zhang L, Zhang X, et al. Synthesis, thermal evolution implantation of Ga in Si and Ge: fluence-dependent retention and optical properties of CuZn alloy nanoparticles in SiO2 and surface morphology. Surface and Interface Analysis, 2008, sequentially implanted with dual ions. Journal of Alloys and 40(11): 1415–1422 Compounds, 2013, 549: 231–237 [297] Stanek S, Nekvindova P, Svecova B, et al. The influence of silver-ion doping using ion implantation on the luminescence [310] Voorhees P W. The theory of Ostwald ripening. Journal of Statistical Physics, 1985, 38(1‒2): 231–252 properties of Er–Yb silicate glasses. Nuclear Instruments & [311] Fan H J, Gösele U, Zacharias M. Formation of nanotubes and Methods in Physics Research Section B: Beam Interactions hollow nanoparticles based on Kirkendall and diffusion with Materials and Atoms, 2016, 371: 350–354 processes: a review. Small, 2007, 3(10): 1660–1671 [298] Fukumi K, Chayahara A, Kadono K, et al. Gold nanoparticles [312] Xu Y H, Wang J P. Direct gas-phase synthesis of ion implanted in glass with enhanced nonlinear optical heterostructured nanoparticles through phase separation and properties. Journal of Applied Physics, 1994, 75(6): 3075–3080 surface segregation. Advanced Materials, 2008, 20(5): 994–999 [299] Yanes A C, del-Castillo J. Enhanced emission via energy [313] Heggen M, Oezaslan M, Houben L, et al. Formation and transfer in RE co-doped SiO2–KYF4 nano-glass-ceramics for analysis of core‒shell fine structures in Pt bimetallic white LEDs. Journal of Alloys and Compounds, 2016, 658: nanoparticle fuel cell electrocatalysts. The Journal of Physical 170–176 Chemistry C, 2012, 116(36): 19073–19083 [300] Wang Y H, Liu F, Cheng H, et al. Third order non-linear [314] Amekura H, Yoshitake M, Plaksin O A, et al. Implantation- optical response of Cu nanoclusters by ion implantation in induced nonequilibrium reaction between Zn ions of 60 keV silicate glass under 532 and 1064 nm laser excitations. Materials Research Innovations, 2014, 18(4): 241–244 and SiO2 target. Applied Physics Letters, 2007, 91(6): 063113 [315] Wang J, Jia G Y, Zhang B, et al. Quasi-two-dimensional Ag [301] Gonella F, Mattei G, Mazzoldi P, et al. Au‒Cu alloy nanoparticle formation in silica by Xe ion irradiation and nanoclusters in silica formed by ion implantation and annealing subsequent Ag ion implantation. Journal of Applied Physics, in reducing or oxidizing atmosphere. Applied Physics Letters, 1999, 75(1): 55–57 2013, 113: 1–8 [316] Peña O, Pal U, Rodríguez-Fernández L, et al. Formation of [302] Cesca T, Calvelli P, Battaglin G, et al. Local-field enhancement Au‒Ag core‒shell nanostructures in silica matrix by sequential effect on the nonlinear optical response of gold‒silver ion implantation. The Journal of Physical Chemistry C, 2009, nanoplanets. Optics Express, 2012, 20(4): 4537–4547 113(6): 2296–2300 [303] Mattei G, Marchi G D, Maurizio C, et al. Chemical- or radiation-assisted selective dealloying in bimetallic nanoclusters. Physical Review Letters, 2003, 90(8): 085502 [304] Liau Z L, Tsaur B Y, Mauer J W. Influence of atomic mixing [317] Torres-Torres C, López-Suárez A, Can-Uc B, et al. Collective optical Kerr effect exhibited by an integrated configuration of silicon quantum dots and gold nanoparticles embedded in ionimplanted silica. Nanotechnology, 2015, 26(29): 295701 and preferential sputtering on depth profiles and interfaces. [318] Yang X C, Li L L, Huang M, et al. In situ synthesis of Ag‒Cu Journal of Vacuum Science and Technology, 1979, 16(2): bimetallic nanoparticles in silicate glass by a two-step ion- 121–127 exchange route. Journal of Non-Crystalline Solids, 2011, [305] Yamada T, Takano A, Sugita K, et al. Effect of dual 357(11‒13): 2306–2308 implantation with Ag and Ni ions on the optical absorption of [319] Obraztsov P A, Nashchekin A V, Nikonorov N V, et al. silica glass. Transactions of the Materials Research Society of Formation of silver nanoparticles on the silicate glass surface Japan, 2020, 45(4): 127–130 after ion exchange. Physics of the Solid State, 2013, 55(6): [306] Cai G X, Ren F, Xiao X H, et al. Morphology control and optical absorption properties of Ag nanoparticles by ion 1272–1278 [320] Garfinkel H M. Ion-exchange equilibria between glass and 78 Front. Mater. Sci. 2022, 16(3): 220607 molten salts. The Journal of Physical Chemistry, 1968, 72(12): 4175–4181 [334] Zhang J, Dong W, Sheng J, et al. Silver nanoclusters formation in ion-exchanged glasses by thermal annealing, UV-laser and [321] Ramaswamy R V, Srivastava R. Ion-exchanged glass waveguides: a review. Journal of Lightwave Technology, 1988, X-ray irradiation. Journal of Crystal Growth, 2008, 310(1): 234–239 [335] Sharma A. Energetic argon beam stimulated growth of 6(6): 984–1000 [322] Crank J. The Mathematics of Diffusion. Oxford, UK: Clarendon, 1956 plasmonic silver nanoparticles in Ag+-exchanged soda glass: a study on the structural, optical, photoluminescence and [323] Kirkwood J G, Oppenheim J. Chemical Thermodynamics. New electrical behavior. Materials Science and Engineering B, 2021, 263: 1–10 York, USA: McGraw Hill, 1961 [324] Zhao J, Zhu J, Yang Z, et al. Selective preparation of Ag [336] Hofmeister H, Thiel S, Dubiel M, et al. Synthesis of nanosized species on photoluminescence of Sm3+ in borosilicate glass via silver particles in ion-exchanged glass by electron beam Ag+‒Na+ ion exchange. Journal of the American Ceramic irradiation. Applied Physics Letters, 1997, 70(13): 1694–1696 [337] Rahman A, Giarola M, Cattaruzza E, et al. Raman Society, 2020, 103(2): 955–964 [325] Zhao J, Yang Z, Yu C, et al. Preparation of ultra-small microspectroscopy investigation of Ag ion-exchanged glass molecule-like Ag nano-clusters in silicate glass based on ion- layers. Journal of Nanoscience and Nanotechnology, 2012, exchange process: energy transfer investigation from molecule- 12(11): 8573–8579 like Ag nano-clusters to Eu3+ ions. Chemical Engineering [326] Karvonen L, Rönn J, Kujala S, et al. High non-resonant thirdorder optical nonlinearity [338] Quaranta A, Rahman A, Mariotto G, et al. Spectroscopic investigation of structural rearrangements in silver ion- Journal, 2018, 341: 175–186 of Ag‒glass nanocomposite fabricated by two-step ion exchange. Optical Materials, 2013, exchanged silicate glasses. The Journal of Physical Chemistry C, 2012, 116(5): 3757–3764 [339] Lenoir M, Grandjean A, Poissonnet S, et al. Quantitation of sulfate solubility in borosilicate glasses using Raman 36(2): 328–332 [327] Mathpal M C, Kumar P, Kumar S, et al. Opacity and plasmonic properties of Ag embedded glass based metamaterials. RSC spectroscopy. Journal of Non-Crystalline Solids, 2009, 355(28‒30): 1468–1473 [340] Mysen B O, Frantz J D. Silicate melts at magmatic Advances, 2015, 5(17): 12555–12562 [328] Kumar P, Mathpal M C, Tripathi A K, et al. Plasmonic temperatures: in-situ structure determination to 1651 °C and resonance of Ag nanoclusters diffused in soda-lime glasses. effect of temperature and bulk composition on the mixing Physical 17(14): behavior of structural units. Contributions to Mineralogy and [329] Simo A, Polte J, Pfänder N, et al. Formation mechanism of [341] Jansen M. Homoatomic d10–d10 interactions: their effects on silver nanoparticles stabilized in glassy matrices. Journal of the structure and chemical and physical properties. Angewandte American Chemical Society, 2012, 134(45): 18824–18833 Chemie International Edition, 1987, 26(11): 1098–1110 Chemistry Chemical Physics, 2015, 8596–8603 Petrology, 1994, 117(1): 1–14 [330] Sheng J. Growth of silver nanoclusters embedded in soda-lime [342] Gonella F, Quaranta A. Stress-induced birefringence in silver- silicate glasses. International Journal of Hydrogen Energy, diffused glass waveguides. Journal of Modern Optics, 1992, 2009, 34(5): 2471–2474 [331] Kumar P, Mathpal M C, Hamad S, et al. Cu nanoclusters in ion exchanged soda-lime glass: study of SPR and nonlinear optical behavior for photonics. Applied Materials Today, 2019, 15: 323–334 [332] Marchi G, Caccavale F, Gonella F, et al. Silver nanoclusters 39(7): 1401–1405 [343] Araujo R. Colorless glasses containing ion-exchanged silver. Applied Optics, 1992, 31(25): 5221–5224 [344] Belharouak I, Weill F, Parent C, et al. Silver particles in glasses of the ‘Ag2O–ZnO–P2O5’ system. Journal of Non-Crystalline Solids, 2001, 293-295: 649–656 formation in ion-exchanged waveguides by annealing in [345] Bromann K, Félix C, Brune H, et al. Controlled deposition of hydrogen atmosphere. Applied Physics A: Materials Science & size-selected silver nanoclusters. Science, 1996, 274(5289): Processing, 1996, 63(4): 403–407 956–958 [333] Rahman A, Mariotto G, Cattaruzza E, et al. Thermal annealing [346] Bastús N G, Comenge J, Puntes V. Kinetically controlled and laser-induced mechanisms in controlling the size and size- seeded growth synthesis of citrate-stabilized gold nanoparticles distribution of silver nanoparticles in Ag+‒Na+ ion-exchanged of up to 200 nm: size focusing versus Ostwald ripening. silicate glasses. Journal of Non-Crystalline Solids, 2021, 563: 120815 Langmuir, 2011, 27(17): 11098–11105 [347] Takesue M, Tomura T, Yamada M, et al. Size of elementary Javier FONSECA. NPs embedded into glass matrices: GNCs clusters and process period in silver nanoparticle formation. Journal of the American Chemical Society, 2011, 133(36): 79 SERS platform for pesticides detection. Sensors and Actuators B: Chemical, 2017, 242: 127–131 [361] Goutaland F, Sow M, Ollier N, et al. Growth of highly 14164–14167 [348] Ye S, Guo Z, Wang H, et al. Evolution of Ag species and concentrated silver nanoparticles and nanoholes in silver- molecular-like Ag cluster sensitized Eu3+ emission in exchanged glass by ultraviolet continuous wave laser exposure. oxyfluoride glass for tunable light emitting. Journal of Alloys Optical Materials Express, 2012, 2(4): 350–357 [362] Goutaland F, Colombier J P, Sow M C, et al. Laser-induced and Compounds, 2016, 685: 891–895 [349] Chen Y, Karvonen L, Säynätjoki A, et al. Ag nanoparticles embedded in glass by two-step ion exchange and their SERS application. Optical Materials Express, 2011, 1(2): 164–172 periodic alignment of Ag nanoparticles in soda-lime glass. Optics Express, 2013, 21(26): 31789–31799 [363] Niry M D, Mostafavi-Amjad J, Khalesifard H R, et al. [350] Berneschi S, Righini G C, Pelli S. Towards a glass new world: Formation of silver nanoparticles inside a soda-lime glass the role of ion-exchange in modern technology. Applied matrix in the presence of a high intensity Ar+ laser beam. Journal of Applied Physics, 2012, 111(3): 033111 Sciences, 2021, 11(10): 4610 [351] Debenedetti P G, Stillinger F H. Supercooled liquids and the glass transition. Nature, 2001, 410(6825): 259–267 [352] Wackerow S, Seifert G, Abdolvand A. Homogenous silverdoped nanocomposite glass. Optical Materials Express, 2011, silver-ion-exchanged glass drawn by infrared nanosecond laser. Nanomaterials, 2020, 10(9): 1849 [365] Wackerow S, Abdolvand A. Laser-assisted one-step fabrication of homogeneous glass–silver composite. Applied Physics A: 1(7): 1224–1231 [353] Sgibnev Y, Asamoah B, Nikonorov N, et al. Tunable photoluminescence of silver molecular clusters formed in Na+‒Ag+ [364] Babich E, Kaasik V, Redkov A, et al. SERS-active pattern in Materials Science & Processing, 2012, 109(1): 45–49 [366] Wackerow S, Abdolvand A. Generation of silver nanoparticles photo-thermo- with controlled size and spatial distribution by pulsed laser refractive glass matrix. Journal of Luminescence, 2020, 226: irradiation of silver ion-doped glass. Optics Express, 2014, ion-exchanged antimony-doped 117411 22(5): 5076–5085 [354] Lumeau J, Zanotto E D. A review of the photo-thermal [367] Zhang J, Dong W, Qiao L, et al. Silver nanocluster formation in mechanism and crystallization of photo-thermo-refractive soda-lime silicate glass by X-ray irradiation and annealing. (PTR) glass. International Materials Reviews, 2017, 62(6): Journal of Crystal Growth, 2007, 305(1): 278–284 348–366 [368] Sonal, Sharma A, Aggarwal S. Optical investigation of soda [355] Nikonorov N V, Panysheva E I, Tunimanova I V, et al. lime glass with buried silver nanoparticles synthesised by ion Influence of glass composition on the refractive index change implantation. Journal of Non-Crystalline Solids, 2018, 485: upon photothermoinduced crystallization. Glass Physics and 57–65 Chemistry, 2001, 27(3): 241–249 [356] Jiao Q, Yu X, Xu X, et al. Relationship between Eu3+ reduction and glass polymeric structure in Al2O3-modified borate glasses under air atmosphere. Journal of Solid State Chemistry, 2013, 202: 65–69 [357] Zhang Q, Liu X, Qiao Y, et al. Reduction of Eu3+ to Eu2+ in Eu-doped high silica glass prepared in air atmosphere. Optical Materials, 2010, 32(3): 427–431 [369] Gangopadhyay P, Magudapathy P, Kesavamoorthy R, et al. Growth of silver nanoclusters embedded in soda glass matrix. Chemical Physics Letters, 2004, 388(4‒6): 416–421 [370] Gangopadhyay P, Magudapathy P, Srivastava S K, et al. Effects of argon ion irradiation on nucleation and growth of silver nanoparticles in a soda-glass matrix. AIP Advances, 2011, 1(3): 032112 [371] Véron O, Blondeau J P, Meneses D D S, et al. Characterization [358] Manikandan D, Mohan S, Magudapathy P, et al. Blue shift of of silver or copper nanoparticles embedded in soda-lime glass plasmon resonance in Cu and Ag ion-exchanged and annealed after a staining process. Surface and Coatings Technology, soda-lime glass: an optical absorption study. Physica B: Condensed Matter, 2003, 325: 86–91 2013, 227: 48–57 [372] Zhang X, Luo W, Wang L J, et al. Third-order nonlinear optical [359] Chervinskii S, Sevriuk V, Reduto I, et al. Formation and 2D- vitreous material derived from mesoporous silica incorporated patterning of silver nanoisland film using thermal poling and with Au nanoparticles. Journal of Materials Chemistry C: out-diffusion from glass. Journal of Applied Physics, 2013, Materials for Optical and Electronic Devices, 2014, 2(34): 114(22): 224301 6966–6970 [360] Tite T, Ollier N, Sow M C, et al. Ag nanoparticles in soda-lime [373] Hirasawa M, Shirakawa H, Hamamura H, et al. Growth glass grown by continuous wave laser irradiation as an efficient mechanism of nanoparticles prepared by radio frequency 80 Front. Mater. Sci. 2022, 16(3): 220607 sputtering. Journal of Applied Physics, 1997, 82(3): 1404–1407 [374] Cattaruzza E, Battaglin G, Canton P, et al. Structural and physical properties of cobalt nanocluster composite glasses. Journal of Non-Crystalline Solids, 2004, 336(2): 148–152 orientation and its contribution to densification during spark plasma sintering. Materials Letters, 2018, 229: 126–129 [387] Mamedov V. Spark plasma sintering as advanced PM sintering method. Powder Metallurgy, 2002, 45(4): 322–328 [375] Chou Y J, Lin S H, Shih C J, et al. The effect of Ag dopants on [388] Marder R, Estournès C, Chevallier G, et al. Plasma in spark the bioactivity and antibacterial properties of one-step plasma sintering of ceramic particle compacts. Scripta synthesized Ag-containing mesoporous bioactive glasses. Journal of Nanoscience and Nanotechnology, 2016, 16(9): 10001–10007 Materialia, 2014, 82: 57–60 [389] Hulbert D M, Anders A, Andersson J, et al. A discussion on the absence of plasma in spark plasma sintering. Scripta Materialia, [376] Palgrave R G, Parkin I P. Aerosol assisted chemical vapor 2009, 60(10): 835–838 to [390] Tokita M. Development of large-size ceramic/metal bulk FGM nanocomposite thin films. Journal of the American Chemical fabricated by spark plasma sintering. Materials Science Forum, deposition using nanoparticle precursors: a route Society, 2006, 128(5): 1587–1597 1999, 308‒311: 83–88 [377] Delgado J, Vilarigues M, Ruivo A, et al. Characterisation of [391] Hu Z Y, Zhang Z H, Cheng X W, et al. A review of multi- medieval yellow silver stained glass from Convento de Cristo physical fields induced phenomena and effects in spark plasma in Tomar, Portugal. Nuclear Instruments & Methods in Physics sintering: fundamentals and applications. Materials & Design, Research Section B: Beam Interactions with Materials and [378] Jiménez J A, Lysenko S, Zhang G, et al. Optical characterization of Ag nanoparticles embedded in aluminophosphate glass. Journal of Electronic Materials, 2007, mesoporous silica with periodic 50 to 300 angstrom pores. Science, 1998, 279(5350): 548–552 [393] Zhang X, Gu S, Zhou B, et al. Solid-state sintering of glasses with optical nonlinearity from mesoporous powders. Journal of 36(7): 812–820 [379] Liu Y F, Liebenberg D H. Electromagnetic radio frequency heating in the pulsed electric current sintering (PECS) process. synthesis sintering/field-assisted of SiC‒TiB2 sintering by technology the American Ceramic Society, 2016, 99(5): 1579–1586 [394] Raven M S. Radio frequency sputtering and the deposition of high-temperature superconductors. Journal of Materials Science MRS Communications, 2017, 7(2): 266–271 [380] Cramer C L, McMurray J W, Lance M J, et al. Reaction-bond composite 2020, 191: 108862 [392] Zhao D, Feng J, Huo Q, et al. Triblock copolymer syntheses of Atoms, 2011, 269(20): 2383–2388 Materials in Electronics, 1994, 5(3): 129–146 plasma [395] Horwitz C M. Radio frequency sputtering — the significance of (SPS/FAST). power input. Journal of Vacuum Science & Technology A, spark Journal of the European Ceramic Society, 2020, 40(4): 988–995 1983, 1(4): 1795–1800 [381] Weston N S, Thomas B, Jackson M. Processing metal powders [396] Mattei G, Battaglin G, Cattaruzza E, et al. Synthesis by co- via field assisted sintering technology (FAST): a critical sputtering of Au–Cu alloy nanoclusters in silica. Journal of review. Materials Science and Technology, 2019, 35(11): Non-Crystalline Solids, 2007, 353(5‒7): 697–702 [397] Cattaruzza E, Battaglin G, Gonella F, et al. Au‒Cu 1306–1328 [382] Gan H, Wang C B, Shen Q, et al. Preparation of La2NiMnO6 nanoparticles in silica glass as composite material for photonic double-perovskite ceramics by plasma activated sintering. applications. Journal of Inorganic Materials, 2019, 34(5): 541–545 1017–1021 [383] Alaniz J E, Dupuy A D, Kodera Y, et al. Effects of applied pressure on the densification rates in current-activated pressureassisted densification (CAPAD) of nanocrystalline materials. [384] Kodera Y, Hardin C L, Garay J E. Transmitting, emitting and controlling light: processing of transparent ceramics using pressure-assisted Surface Science, 2007, 254(4): [398] Okuyama K. Preparation of micro-controlled particles using aerosol process. Journal of Aerosol Science, 1991, 22: S7–S10 [399] Messing G L, Zhang S C, Jayanthi G V. Ceramic powder synthesis by spray pyrolysis. Journal of the American Ceramic Scripta Materialia, 2014, 92: 7–10 current-activated Applied densification. Scripta Materialia, 2013, 69(2): 149–154 [385] Shen Z, Johnsson M, Zhao Z, et al. Spark plasma sintering of alumina. Journal of the American Ceramic Society, 2002, 85(8): 1921–1927 [386] Deng S, Li R, Yuan T, et al. Effect of electric current on crystal Society, 1993, 76(11): 2707–2726 [400] El-Kady A M, Ali A F, Rizk R A, et al. Synthesis, characterization and microbiological response of silver doped bioactive glass nanoparticles. Ceramics International, 2012, 38(1): 177–188 [401] Hong Y L, Liu Z, Wang L, et al. Chemical vapor deposition of layered two-dimensional MoSi2N4 materials. Science, 2020, 369(6504): 670–674 Javier FONSECA. NPs embedded into glass matrices: GNCs 81 [402] Cai Z, Liu B, Zou X, et al. Chemical vapor deposition growth [414] Ono M, Hata M, Tsunekawa M, et al. Ultrafast and energy- and applications of two-dimensional materials and their efficient all-optical switching with graphene-loaded deep- heterostructure. Chemical Reviews, 2018, 118(13): 6091–6133 subwavelength plasmonic waveguides. Nature Photonics, 2020, [403] Pu J, Tang L, Li C, et al. Chemical vapor deposition growth of few-layer graphene for transparent conductive films. RSC Advances, 2015, 5(55): 44142–44148 14(1): 37–43 [415] Yang Y, Kelley K, Sachet E, et al. Femtosecond optical polarization switching using a cadmium oxide-based perfect [404] Lozovoy K A, Korotaev A G, Kokhanenko A P, et al. Kinetics absorber. Nature Photonics, 2017, 11(6): 390–395 of epitaxial formation of nanostructures by Frank‒van der [416] Ren M, Jia B, Ou J Y, et al. Nanostructured plasmonic medium Merwe, Volmer‒Weber and Stranski‒Krastanow growth for terahertz bandwidth all-optical switching. Advanced modes. Surface and Coatings Technology, 2020, 384: 125289 [405] Zinke-Allmang M. Phase separation on solid surfaces: nucleation, coarsening and coalescence kinetics. Thin Solid Films, 1999, 346(1‒2): 1–68 Materials, 2011, 23(46): 5540–5544 [417] Pilot R, Signorini R, Durante C, et al. A review on surfaceenhanced Raman scattering. Biosensors, 2019, 9(2): 57 [418] Kolobkova E, Kuznetsova M S, Nikonorov N. Ag/Na ion [406] Ertorer E, Avery J C, Pavelka L C, et al. Surface-immobilized exchange in fluorophosphate glasses and formation of Ag gold nanoparticles by organometallic CVD on amine- nanoparticles in the bulk and on the surface of the glass. ACS terminated glass surfaces. Chemical Vapor Deposition, 2013, 19(10‒12): 338–346 Applied Nano Materials, 2019, 2(11): 6928–6938 [419] Takada K. Progress and prospective of solid-state lithium [407] Hamilton J A, Pugh T, Johnson A L, et al. Cobalt(I) olefin batteries. Acta Materialia, 2013, 61(3): 759–770 complexes: precursors for metal-organic chemical vapor [420] Thangadurai V, Weppner W. Recent progress in solid oxide deposition of high purity cobalt metal thin films. Inorganic and lithium ion conducting electrolytes research. Ionics, 2006, Chemistry, 2016, 55(14): 7141–7151 12(1): 81–92 [408] Palgrave R G, Parkin I P. Aerosol assisted chemical vapor [421] Kim J G, Son B, Mukherjee S, et al. A review of lithium and deposition of gold and nanocomposite thin films from hydrogen non-lithium based solid state batteries. Journal of Power tetrachloroaurate(III). Chemistry of Materials, 2007, 19(19): 4639–4647 [409] Ashraf S, Blackman C S, Hyett G, et al. Aerosol assisted chemical vapour deposition of MoO3 and MoO2 thin films on glass from Sources, 2015, 282: 299–322 [422] Takada K, Aotani N, Iwamoto K, et al. Solid state lithium molybdenum polyoxometallate battery with oxysulfide glass. Solid State Ionics, 1996, 86‒88: 877–882 precursors; [423] Liu L, Zhao F, Chen X, et al. Local delivery of FTY720 in thermophoresis and gas phase nanoparticle formation. Journal mesoporous bioactive glass improves bone regeneration by of Materials Chemistry, 2006, 16(35): 3575–3582 synergistically [410] Talbot L, Cheng R K, Schefer R W, et al. Thermophoresis of particles in a heated boundary layer. Journal of Fluid Mechanics, 1980, 101(4): 737–758 immunomodulating osteogenesis and osteoclastogenesis. Journal of Materials Chemistry B, 2020, 8(28): 6148–6158 [424] Shoaib M, Saeed A, Rahman M S U, et al. Mesoporous nano- [411] Ponja S D, Williamson B A D, Sathasivam S, et al. Enhanced bioglass designed for the release of imatinib and in vitro electrical properties of antimony doped tin oxide thin films inhibitory effects on cancer cells. Materials Science and deposited via aerosol assisted chemical vapour deposition. Journal of Materials Chemistry C: Materials for Optical and Electronic Devices, 2018, 6(27): 7257–7266 [412] Gardecka A J, Goh G K L, Sankar G, et al. On the nature of niobium substitution in niobium doped titania thin films by Engineering C, 2017, 77: 725–730 [425] Tabia Z, El Mabrouk K, Bricha M, et al. Mesoporous bioactive glass nanoparticles doped with magnesium: drug delivery and acellular in vitro bioactivity. RSC Advances, 2019, 9(22): 12232–12246 AACVD and its impact on electrical and optical properties. [426] Shuai C, Xu Y, Feng P, et al. Antibacterial polymer scaffold Journal of Materials Chemistry A: Materials for Energy and based on mesoporous bioactive glass loaded with in situ grown Sustainability, 2015, 3(34): 17755–17762 silver. Chemical Engineering Journal, 2019, 374: 304–315 [413] Walters G, Parkin I P. Aerosol assisted chemical vapour [427] Fonseca J, Choi S. Rational synthesis of a hierarchical deposition of ZnO films on glass with noble metal and p-type supramolecular porous material created via self-assembly of dopants; use of dopants to influence preferred orientation. metal-organic framework nanosheets. Inorganic Chemistry, Applied Surface Science, 2009, 255(13‒14): 6555–6560 2020, 59(6): 3983–3992