Next Article in Journal
Carbon Dioxide Reduction with Hydrogen on Fe, Co Supported Alumina and Carbon Catalysts under Supercritical Conditions
Next Article in Special Issue
Microwave-Assisted Sequential One-Pot Synthesis of 8-Substituted Pyrazolo[1,5-a][1,3,5]triazines
Previous Article in Journal
Multiconfiguration Pair-Density Functional Theory for Transition Metal Silicide Bond Dissociation Energies, Bond Lengths, and State Orderings
Previous Article in Special Issue
Synthesis of Isothiocyanates Using DMT/NMM/TsO as a New Desulfurization Reagent
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Microwave-Assisted Preparation of Luminescent Inorganic Materials: A Fast Route to Light Conversion and Storage Phosphors

1
Institute of Physics, University of São Paulo, São Paulo BR-05508-900, SP, Brazil
2
The Molecular Foundry, Lawrence Berkeley National Laboratory, Berkeley, CA 94720, USA
3
Institute of Chemistry, University of São Paulo, São Paulo BR-05508-000, SP, Brazil
4
Nuclear and Energy Research Institute, São Paulo BR-05508-000, SP, Brazil
*
Author to whom correspondence should be addressed.
Molecules 2021, 26(10), 2882; https://doi.org/10.3390/molecules26102882
Submission received: 26 March 2021 / Revised: 7 May 2021 / Accepted: 9 May 2021 / Published: 13 May 2021
(This article belongs to the Special Issue The Application of Microwave Technology in Chemistry)

Abstract

:
Luminescent inorganic materials are used in several technological applications such as light-emitting displays, white LEDs for illumination, bioimaging, and photodynamic therapy. Usually, inorganic phosphors (e.g., complex oxides, silicates) need high temperatures and, in some cases, specific atmospheres to be formed or to obtain a homogeneous composition. Low ionic diffusion and high melting points of the precursors lead to long processing times in these solid-state syntheses with a cost in energy consumption when conventional heating methods are applied. Microwave-assisted synthesis relies on selective, volumetric heating attributed to the electromagnetic radiation interaction with the matter. The microwave heating allows for rapid heating rates and small temperature gradients yielding homogeneous, well-formed materials swiftly. Luminescent inorganic materials can benefit significantly from the microwave-assisted synthesis for high homogeneity, diverse morphology, and rapid screening of different compositions. The rapid screening allows for fast material investigation, whereas the benefits of enhanced homogeneity include improvement in the optical properties such as quantum yields and storage capacity.

Graphical Abstract

1. Introduction

1.1. Luminescent Materials: Optical Properties and Applications

Luminescent materials, also known as phosphors, usually refers to inorganic materials that emit light after external physical stimuli, such as UV-VIS, IR-Lasers, X-rays, γ-rays, and e-beams [1]. These materials are composed of a host lattice with intentionally added optically active impurities (dopants). The dopants act as luminescent centers, converting the incident excitation into a characteristic light emission that depends on the dopant’s identity [2]. The most common dopants for luminescent materials are the f-metals, such as the trivalent rare-earths (RE3+: Pr, Nd, Eu, Tb, Dy, Er), that emit light after optical transitions within the f-electron manifold [2]. The f-electrons are strongly shielded from the chemical environment and retain atomic characteristics, yielding sharp emission lines with wavelengths nearly independent of the host’s composition. The f-f electronic transitions are parity forbidden and, thus, the emission decay times are long (μs-ms).
Some rare earth ions (e.g., Ce3+, and Eu2+) behave as crystal-field sensitive ions, having intense broad emission bands with emission color dependent on the host’s composition. In these cases, the emission is governed by parity-allowed 4f-5d transitions, yielding intense emission bands with fast decay times (ns-μs). Figure 1 provides the comparison between the f-f and f-d transitions for the Eu3+ and Eu2+ ions.
Various transition metals also act as luminescent centers, e.g., Mn4+-doped K2SiF6 [5], and Cr3+-doped ZnGa2O4 [6]. In these cases, parity-forbidden d-d electronic transition is responsible for the emission. The d-orbitals broadly interact with the chemical environment leading to strong crystal field splitting according to the composition of the materials. Thus, most of the transition metals’ absorption and emission bands are broad. However, sharp emissions can also be observed, such as Mn4+-doped K3ScF6 red phosphors [7].
Luminescent materials can act as light converters due to different photophysical processes, such as downconversion (or quantum cutting) [8,9,10], upconversion [11,12,13], and down-shifting [14,15] (Figure 2A–C). Downshifting is the most conventional photoluminescence mechanism, where high-energy photons are converted into low-energy photons. Ideally, all excited photons are converted to emitted photons without losses in the optical process, obtaining a quantum yield of 100% [16,17,18,19]. On the other hand, the first experimental evidence for quantum yields higher than 100% was reported for YF3:Pr3+ material [17,20]. This optical phenomenon proposed by Dexter consists of the high-energy photons splitting into two or more lower-energy photons. This mechanism, called downconversion or quantum cutting, involves single or multiple ions (e.g., in YF3:Pr3+ [17,20] or YPO4:Yb3+, Tb3+ [10], respectively).
Upconversion (UC) is a nonlinear optical phenomenon that converts two or more low-energy photons to a higher-energy photon [13]. This optical process, known as anti-Stokes emission, was theoretically proposed by the physicist Nicolaas Bloembergen in 1959 [21]. The first experimental result, reported by François Auzel in 1966, consisted of the energy transfer from Yb3+ to Er3+ and Tm3+ ions [13]. The UC process has since inspired light conversion mainly from near-infrared (NIR) to UV-visible-NIR light, suitable for bioimaging, solar cells, anticounterfeiting, photocatalysis, and photodynamic therapy [22].
Persistent luminescence (PeL) materials continue to emit light after removing the excitation source due to stored energy in the structural defects that act as charge carrier traps [23,24,25]. After the removal of the excitation, the material absorbs available thermal (kT) or optical (hv) energy to release the trapped charge carriers that recombine with the emitting centers. The radiative decay of the excited states emits a photon with a characteristic wavelength. The most efficient PeL phosphors are dominated by Eu2+-doped inorganic materials, exhibiting 24 to 48 h of persistent emission after few minutes of excitation in the UV-VIS range [25]. However, plenty of examples for d- and f-metals are available, and extensive reviews on the topic have been published [23,24].
Luminescent materials have several applications in technology, telecommunications, and health sciences, and are the subject of extensive research. Specific applications can be in light-emitting diodes [26,27,28], laser [29], thermometry [30], thermoluminescence dating [31,32], dosimetry [33], bioimaging [6,34], and photodynamic therapy [35,36,37].
For instance, phosphor converters for LEDs (pc-LEDs) work as light converters where an excitation source, e.g., blue emitting-LEDs, excites the luminescent material that emits broadband in the yellow region. This kind of phosphor has defined characteristics, such as high quantum yields and fast excited-state decay times. The most efficient light converters for pc-LED are inorganic hosts doped mainly with Ce3+ and Eu2+ [27,38,39,40]. The parity-allowed (f–d) electronic transition in Ce3+ and Eu2+ ions results in high-intensity emissions with short excited-state lifetimes (μs-ns). However, other examples can be easily found in trivalent rare earths, Eu3+, Tb3+, Pr3+, Dy3+, and d-metals (Mn4+) doped materials [5,7,41,42,43,44]. PeL materials are in evidence nowadays due to vast and unexplored bioimaging applications [37], theranostics [45], and photocatalysis [46]. When it comes to application, persistent phosphors must have high quantum yield and high storage capacity [47,48].

1.2. Concepts and Mechanisms of Microwave Heating

Microwaves (MW) are electromagnetic radiation in the frequency interval of 300 MHz (1.24 μeV; 1 m) to 300 GHz (1.24 meV; 1 mm). Radiation in this range is mainly used in wireless telecommunication protocols, such as Bluetooth and wi-fi modems [49]. A narrow part of the microwave spectrum is used for industrial, domestic, scientific, and medical purposes (ISM bands) [50]. The industrial and domestic uses of microwaves come from the ability to heat certain materials depending on their dielectric properties [49,51]. For instance, the radiation in a domestic microwave oven (2.45 GHz; λ: 12.2 cm) interacts with the water contained in the food, heating it swiftly due to the intrinsic dielectric properties of the water molecules. Further, a precise tuning of microwave radiation parameters enables microwave heating of different materials, such as metal oxides [52,53], carbon [49,54,55], and organic solvents [56].
Microwave heating properties differ significantly from conventional heating ones. In conventional heating, the external source, such as a resistive furnace, heats the material’s surface, leading to a large temperature gradient compared to the material’s core. Instead, dielectric heating occurs volumetrically, meaning that the material irradiated by the microwaves is heated locally, leading to a small temperature gradient [56] (Figure 3). Microwave-assisted heating rates are drastically steeper than conventional heating because of the absence of heat transfer from an external source.
The combination of these properties can be highly beneficial to the synthesis of inorganic materials using microwave-assisted approaches. For instance, solid-state methods benefit significantly from dielectric heating, shortening the synthesis time from days and hours to tens of minutes. In addition, it can yield more uniform, homogeneous materials [57].
The principles of microwave heating are based on the material’s response to the electromagnetic fields’ alternating nature. Bhattacharya et al. [49] discussed in detail the microwave heating mechanism in several materials. The relevant parameters for the heating include the dielectric constant (ε′) and dielectric loss (ε′′). The parameters can be interpreted as the ability of the material to store (ε′) and convert (ε′′) the electric energy from the microwave irradiation.
The main mechanism of dielectric heating is attributed to the polarization losses that comprise dipolar, interfacial, ionic, and electronic losses [49,51]. All these contributions are related to the intrinsic properties of the materials. Thus, a corrected assessment of the microwave interaction with a particular material must include a detailed analysis of its dielectric properties. Additionally, if the material to be irradiated has high conductivity, then the conduction loss enters the summation in the dielectric loss constant (ε′′). Magnetic contributions must be considered when the material presents a magnetic moment. Most luminescent materials are not intrinsically magnetic, making it possible to simplify the equation to only the electric contribution.
The interaction of microwaves with matter can be understood as how much a microwave can penetrate in a specific material. The penetration depth of a microwave within materials interacting only with the electric field (μ″r = 0) can be expressed by Equation (1):
d p = c 2 π f ε r μ r [ ( 1 + tan 2 δ e ) 1 ] 1 / 2
where f is the frequency of the microwave, (ε′r) is the relative dielectric constant, and (ε′′r) is the relative dielectric loss. The loss tangent ( tan δ e ) is an important parameter that denotes the dissipation of the electrical energy quantitatively due to multiple physical processes [58], and can be defined by Equation (2):
tan δ e = ε r ε r
The dp parameter will then classify the materials’ interaction with the microwave in three different classes, (i) opaque or reflector, (ii) transparent, and (iii) absorber [49]. If the penetration depth is shallow (micrometer range), the microwave is reflected by the material’s surface. Typical examples of microwave reflectors are metals like Cu (dp = 1.3 μm) and Al (dp = 1.7 μm). Alternatively, if the penetration depth is too large, then the microwave goes through the material without any interaction being transmitted. Materials like alumina, borosilicate glasses, and zirconia are excellent examples of microwave-transparent materials, with penetration depths larger than 10 m.
Neither of these situations is ideal for heating the materials and therefore performing microwave-assisted synthesis. To interact with microwaves, the radiation’s penetration depth must be optimal to dissipate the electrical energy efficiently. Absorber materials have penetration depths in the order of centimeters, such as SiC (dp = 1.9 cm), water (3.0 cm), and carbon black (5.7 cm) [59,60,61]. The tangent loss of the materials in terms of reflector, transparent and absorber have typical values of tan δ e   < 0.01 (transparent), >0.1 (absorbers), and >>1 (reflectors).
The fact that the dielectric properties of the materials can vary with temperature increases possibilities of interacting with materials by changing their reactional medium temperature. For instance, yttria-stabilized zirconia is microwave-transparent up to 200 °C ( tan δ e = 0.001 ; d p = 5.68   m ), suffering a transition to an absorber material at 600 °C ( tan δ e = 0.15 ; d p = 0.039   m ). The principle of intrinsic tangent loss modulated by the materials’ temperature is beneficial for synthetic purposes. A synthetic precursor consisting of a mixture of low-lossy materials (transparent) can be in close contact with a high-lossy (absorber) material that will initially interact with the microwave, generating local heating. The heat transfer to the low-lossy material can inflict a change in the tangent loss, transitioning from transparent to absorber. This type of heating can be denominated as hybrid microwave heating, and it is the most-used method in microwave-assisted solid-state (MASS) synthesis [49,51,62,63].

1.3. Properties of Microwave Heating

Microwave-assisted heating can be characterized by a combination of unique properties offering faster and more efficient heating, often leading to more crystalline, homogeneous materials [64]. Firstly, the direct energy conversion into heat leads to a smaller temperature gradient throughout the precursors. In conventional resistive heating, the large temperature gradients make the reactions sluggish, taking a long time to reach equilibrium and generate the products. Dielectric heating works in the opposite manner, heating the materials from the inside to the outside. Dielectric heating leads to minimal energy consumption and faster reaction rates [65]. The thermal efficiency is an appealing point when considering laboratory work for sample screening and compositional tailoring. High throughput can be obtained with specialized setups, allowing the synthesis of tens of samples in few minutes. The fast screening in compositional tailoring can be beneficial for luminescent materials with a strong correlation of the optical property with the composition and structure.
Dielectric heating is also volumetric by nature, meaning that the heating is uniform throughout the sample, in theory. In actual samples, the volumetric heating is not perfect due to inhomogeneity in the precursors and temperature dependence of the dielectric properties [66]. However, compared with conventional heating, the MASS method has more uniform heating and, therefore, a more homogeneous composition of the final products. For luminescent materials, this property is of utmost importance once the homogeneous dispersion of activator and sensitizer ions in the material’s host leads to higher quantum yields and increased brightness [67]. Homogeneous luminescent materials offer a more-defined local structure that can lead to a precise control of the emitting color of crystal field-sensitive ions, such as Eu2+ [57].
The ionic diffusion limit encountered in conventional heating methods is also a significant problem in solid-state synthesis [64,68]. Usually, the diffusion is exceptionally slow if the temperature of the synthesis is lower than two thirds of the precursor’s melting point (Tamann’s rule) [69], which translates to temperatures usually higher than 1000 °C. The microwave radiation effect on ionic diffusion in bulk inorganic materials was investigated by Whittaker et al. [70]. The use of polarized MW irradiation combined with x-ray fluorescence technique has shown a preferential ionic migration in the polarized axis of the radiation (Figure 4a–b).
The increased diffusion is related to increased reaction rates in microwave-heated synthesis in comparison with resistive furnaces. Vanetsev et al. [71]. studied the kinetics of NiFe2O4 formation by conventional and microwave-assisted solid-state reaction. Besides the faster reaction rates, microwave irradiation also promotes changes in the rate-controlling stage of the synthesis. By removing the diffusion hindrances due to non-thermal ponderomotive forces, the rate-controlling step of the synthesis changes from ion-diffusion to the interfacial chemical reaction step (Figure 4c) [71].
Instantaneous heating is also an essential feature of dielectric heating, swiftly converting the MW energy into heat. Correspondingly, once the microwave irradiation is ceased, the heating stops immediately, causing the reaction to quench instantaneously. This could lead to metastable phases, otherwise impossible to synthesize using conventional methods. The temperature quenching can be greatly beneficial for persistent luminescence phosphors because of the defect structure caused by rapid heating and cooling rates.
Once the dielectric constant and tangent loss vary with the material composition and structure, microwave irradiation allows for selective heating of the reactants [64]. This effect can find incredible applications on heating specific sites in supported metal catalysts [72]. The synthesis of chalcogenides greatly benefits the selective heating, allowing the reaction to occur before the sulfur evaporation [73].
The properties of microwave-assisted heating are summarized in Figure 5. The combination of features allows for the synthesis of highly efficient luminescent materials with singular properties, as will be discussed in detail in the next sections.

2. Methodologies of Microwave-Assisted Synthesis

There are several approaches to obtaining luminescent materials using microwave-assisted methodologies. However, microwave-assisted syntheses needs a detailed design of the experimental apparatus due to many dielectric heating parameters. The synthesis setup needs to be modified depending on the application to achieve optimum performance and reproducible results. Here, we present some details on the most common methods to obtain solid inorganic luminescent materials using microwave-assisted heating. Each one has advantages and disadvantages, and the materials’ application should be considered when choosing the method as will be discussed further in next sections.

2.1. Microwave-Assisted Solid-State (MASS) Synthesis

Although there are some commercial setups for MASS synthesis, most of the literature relies on the usage of a domestic microwave oven (DMO) [49,54,55,63,74,75]. This is probably because of two main reasons: (i) economic viability: DMOs are inexpensive equipment, and (ii) DMOs are easily customizable to meet the needs of many methods.
The usage of DMOs comes with the price of several safety issues that need to be addressed carefully. In a customized DMO setup, the researcher needs to avoid radiation leakage, harmful to human body tissues at the frequency of 2.45 GHz. Additionally, DMO setups should be placed in a fume hood to avoid inhalation of toxic gasses from reactions. Effective thermal insulation of the reactional vessel is indispensable to prevent damage to the microwave oven.
To achieve optimum reproducibility, one needs to be consistent with the parameters of the synthesis, with a carefully chosen irradiation program. Crucial synthesis parameters are the crucible sizes, amount of starting materials, nature and mass of the susceptor, position inside the microwave, radiation frequency, and irradiation power levels [63].
A typical setup for microwave-assisted solid-state synthesis is depicted in Figure 6a,b. The setup for MASS (Figure 6b) synthesis consists of three parts [76,77]:
(i) Reaction vessel: crucible where the precursor powder will be placed, and the reaction occurs. Usually, the crucibles are made of alumina (Al2O3) due to its high thermal and chemical stability. For specific purposes other crucibles can be used, such as zirconia or graphite. The ideal size of the crucibles is up for debate, but crucibles varying from 3 to 10 mL in volume are the most common.
(ii) Susceptor: material that absorbs 2.45 GHz microwaves efficiently at room temperature. Due to the fixed frequency of the radiation in DMOs, few oxides directly interact with the microwave. Since a significant number of luminescent materials use oxides as starting materials, most of the works in literature use the hybrid heating approach, where a susceptor (e.g., carbon) is used to initiate the heating (Table 1). The susceptor also has a secondary effect of generating a local atmosphere. The composition of the local atmosphere must be considered when synthesizing the materials. For example, a mildly reducing atmosphere is obtained when using carbon as a susceptor due to its incomplete burn that generates CO gas.
(iii) Thermal insulation: Insulating blocks where the reaction and susceptor crucibles will be placed to insulate extreme temperatures of the solid-state synthesis. Usually aluminosilicate bricks are used to host the reaction system (precursors + susceptors). Alumina foam and glass wool are also reported in the literature. The criteria to choose the correct system is the maximum thermal insulation allied to the high MW transparency.
The three-part system is inserted in a DMO and irradiated using time vs. power level program usually defined by measuring the temperature increase rate. Although temperature measurement is not an easy task, it is of utmost importance to avoid underreacted samples or decomposition by excessive temperatures [76]. A pyrometer is a convenient tool to measure the instantaneous temperature of the crucible. However, the recorded temperature is just of the material’s surface. The temperature inside the susceptor should be considerably higher because the microwave penetration promotes heating initially below the surface of the material [78]. The same effect can be expected for the sample, where the core is hotter than the surface. Despite the difference in temperature between the core and surface, the use of hybrid heating using a susceptor helps obtain a more homogeneous heating of the samples compared to direct heating [78].
Carbon is the most-used susceptor and can absorb microwave radiation from DMOs reaching rapidly high temperatures (Table 1). Additionally, carbon undergoes through an incomplete burn at high temperatures, generating CO reducing gas. The local, reducing atmosphere can be beneficial if the luminescent material to be prepared needs to reduce activator ions’ initial oxidation state. For instance, Eu2+ doped in Sr2MgSi2O7 phosphors, which need to be reduced from the trivalent state in the precursor, e.g., Eu2O3 [57]. Tb4O7 is a common precursor for Tb-doped phosphors prepared by the solid-state method, sometimes needing H2 gas to be reduced from Tb4+ to Tb3+ [82]. With the MASS method, it is possible to achieve Lu2O3:Tb3+ in a pure trivalent state, in 20 min, using activated carbon as susceptor (Figure 7) [62]. When preparing Y3Al5O12:Ce3+, the pure tetravalent form of cerium oxide, CeO2, is reduced to Ce3+ during the MASS synthesis in a single step of ca. 20 min [83].
Silicon carbide, SiC, is also a good option for susceptors because of its high-temperature stability (2830 °C). SiC is common in commercial microwave ashing furnaces that use SiC plates or rods to convert the MW radiation into heat (Figure 6a) [84]. Despite a slow initial heating rate, SiC has a monotonic increase of temperature, from 25 to 600 °C, working as a steady microwave heating element. After 600 °C, the microwave heating rate increases dramatically, reaching 1800 °C in few minutes [80]. CuO has some usage as a reactant‑susceptor, used as a reactant in the MASS synthesis of YBa2Cu3O7−x superconductors [53].

2.2. Microwave-Assisted Combustion (MAC)

Combustion relies on rapid oxidation that generates heat or as a sequence of exothermic chemical reactions between a fuel and an oxidant that is accompanied by the production of heat and the conversion of chemical species [85,86,87,88]. In the volume reaction model, also referred to as thermal explosion, the entire sample is uniformly heated to the ignition temperature, and the combustion reaction coincides in all parts of the sample.
The conventional combustion synthesis uses a certain amount of oxidizer (e.g., metal nitrates) and fuel (e.g., urea) to produce an aqueous solution that, upon heat, ignites the combustion [87,88]. The microwave-assisted method works in the same way, but the formed solution needs to absorb the microwave radiation to initiate the heat. Several solvents (Table 2) have a suitable dielectric tangent loss to interact with the microwave electric field and convert it into heat.
The combustion synthesis has many advantages such as high reaction temperatures with relatively low preheating temperatures, high heating rates, short duration of reactions, relatively simple equipment, materials with porous morphology, and varied microstructure [89]. The microwave-assisted combustion makes the synthesis even faster, in the order of tenths of seconds (Figure 8).
The experimental setup for the MAC method can be obtained with a modified DMO. Nevertheless, attention to the safety issues must be thoroughly considered due to physical and chemical hazards involved in the combustion flames and toxic fumes. Moreover, microwave leakage from the modified setup should be avoided at all costs. An example of the synthesis configuration used for microwave-assisted combustion synthesis is depicted in Figure 8.

2.3. Microwave-Assisted Hydrothermal (MAH)

Hydrothermal synthesis refers to a crystallization technique of compounds under high-temperature aqueous solutions and at high vapor pressures. The crystal growth is performed in an apparatus consisting of a steel pressure vessel called an autoclave. A temperature gradient is maintained between the opposite ends of the chamber where the hotter end dissolves the ions. At the cooler end the crystals are formed. Advantages of the hydrothermal method include the ability to create crystalline phases which are not stable at the water boiling point.
The vessel barrier in the hydrothermal apparatus makes the synthesis in a resistive oven long and energetically inefficient. On the other hand, the microwave-assisted hydrothermal (MAH) method is an advantageous and powerful method to synthesize inorganic materials in a noticeably short time. The microwave radiation interacts directly with the water heating the solution in a short time. It makes the MAH synthesis highly efficient regarding energy conversion. The pressured solution allied to high temperatures allows crystallization of particles in the nanoscale as well. The vast number of compounds soluble in water, temperature modulation and pressure control allow researchers obtain nanoparticles with different composition, crystalline structures and morphologic shapes (rings, crystals, rods, spheres, cubes, urchin) [91].
Although there are examples of MAH synthesis using a modified DMO, it is advised to use a specialized setup due to the risk of explosion of the pressurized vessel [92]. Luckily, high-quality commercial setups are available. Among the advantages of MAH synthesis, the most surprising is obtaining nanoparticles with a narrow particle size distribution and varied morphologies (Figure 9). This is possible due to the homogeneous and fast dielectric heating by the microwave radiation.

2.4. Microwave-Assisted Sol-Gel (MASG)

Sol-gel synthesis combines two synthetical methodologies. In the first step, a polymeric 3D-framework is obtained by using organometallics or silane precursors. These precursors undergo hydrolysis and condensation process to form a gel, that contains evenly dispersed ions in the polymeric matrix. In a second step, to obtain the final material the gel precursor is thermally treated to remove carbon moieties. In the conventional sol-gel method, the heat treatment is performed using a resistive furnace. The MASG method utilizes microwave-assisted heating to thermally treat the gel precursors.
Birkel et al. [93]. obtained Pd-substituted LnFeO3 (Ln: Y, and La) using the MASG method. The prepared gel using citric acid and metal acetylacetonates was dried and then ground to obtain a powder that was subsequently heated in a microwave using the MASS setup. The noticeably short heating times in microwave reactions allowed the elucidation of the phase formation process showing that the reaction for forming LaFeO3 is completed after 90 s. In comparison with the furnace-based sol–gel or combustion methods, the MASG method has allowed the easy and reproducible tuning of material properties such as crystallite size and surface area, which are important parameters for photocatalytic activity.
Near-infrared (NIR) persistent luminescence nanoparticles are widely investigated for bioimaging and theranostics. However, if the particle’s hydrodynamic radius is bigger than 100 nm, the particles are sequestrated by the reticuloendothelial system (RES). Non-aqueous MASG synthesis can be an efficient route to obtain ultra-small nanoparticles (average size: 6 nm) of ZnGa2O4:Cr3+ persistent phosphors [94]. The ultra-small nanoparticle size reduces the RES clearance, enhances the blood circulation time, and allows widespread organ distribution. The non-aqueous benzyl alcohol route has been established as a powerful technique that takes place at a moderate temperature and pressure to obtain exceedingly small metal oxide nanoparticles with high crystallinity, purity, and reproducibility. The combination with the microwave heating offers a reduction of the reaction time (30 min) in comparison to traditional heating in an autoclave (48 h) [95]. One special advantage of this technique is that the one-pot process does not need further high temperature treatment, which generally leads to excessive particle growing.

2.5. Other Preparation Methods

Coupled methodologies with microwave-assisted heating are an interesting alternative for synthesizing luminescent materials. Bi et al. [96] used a coupled microwave-ultrasound-assisted method to obtain luminescent Y2O3:Eu3+ materials. The method consists of the irradiation of a Y(NO3)3 water/ethanol solution using ultrasound (100 W) and microwave (150 W) for one hour. The method allows for nanoparticle formation (<100 nm) due to the local heating leading to fast crystallization and the ultrasound avoiding the excessive coalition of smaller particles. In this case, the obtained luminescent materials present high quantum yields of emission (ca. 80%).
The co-precipitation method consists of precipitate substances normally soluble in the precursor solution. In this method the mixing of the solution containing the soluble cations and anions forces the parts to interact, inducing nucleation and growth. After the particles surpass the critical mass the precipitation of the materials is observed. Usually aqueous solutions are used; however, it is possible to obtain precipitates in other polar solvents. Thongtem et al. [97] used a microwave-assisted co-precipitation method to obtain SrMO4 (M: W, and Mo) nanocrystals. The crystallization of the materials was performed by irradiating an ethylene glycol (EG) dispersion of the NaMO4 and Sr(NO3)2 salts. The effect of the EG solvent in the synthesis is not discussed, but the mildly chelating effect might play a role in retarding the precipitation and crystallization. Moreover, EG is an excellent microwave absorber, leading to a homogeneous heating profile. Due to the controlled crystallization in this method, they were able to produce small nanoparticles (~25 nm) with a narrow size distribution.
The comparison methods of main microwave-assisted methods is summarized in Table 3.

3. Inorganic Luminescent Materials Obtained by Microwave-Assisted Methods

Inorganic luminescent materials can be obtained by various methods, such as conventional solid-state (CSS) synthesis, combustion synthesis, or sol–gel methods. However, the CSS method is the most widely used due to the combination of the easy processing and few steps that grants high reproducibility. In the CSS method, a high-temperature, direct reaction is expected where the solid reactants, with different grain sizes and randomly oriented surfaces, must react through ionic diffusion. The drawback of the solid-state reactions is the limited flux of diffusing species of precursors, impeding crystal growth and forming the final products [57]. The combination of low ionic diffusion with large temperature gradient during the resistive heating translates to long synthesis’ time at high temperatures (>1000 °C) to avoid inhomogeneous materials with poor optical properties. Thus, increasing the ionic diffusion and reducing the temperature gradient can be an alternative method of obtaining efficient inorganic luminescent materials swiftly. In the following sections, we review several classes of luminescent materials obtained by microwave-assisted methodologies.

3.1. Oxides

Several luminescent materials are based on oxides doped with activator ions, e.g., Y2O3:Eu3+ [96], Lu2O3:Tb3+ [101]. Usually, oxide luminescent materials are prepared from oxide precursors in a solid-state synthesis. In addition, carbonate precursors are commonly used with the same intent to obtain the respective oxide after the decomposition of carbon moieties at moderate temperatures. Since most of the metal oxides are low-lossy materials, having considerable penetration depth, they act as microwave-transparent materials. Thus, hybrid heating is used to achieve the needed high temperature for the synthesis.
Pedroso et al. [62] used the MASS method to obtain the Lu2O3:RE3+ (RE: Pr, Eu, and Tb) persistent luminescent materials using a modified DMO setup. Rare-earth doped lutetia is a known family of materials with long persistent luminescence decay, especially the Lu2O3:Tb3+, Ca2+ [82]. Lu2O3-based materials require rather extreme preparation conditions in solid-state synthesis needing high temperature (~1700 °C) in high vacuum or H2-N2 gas flow [102]. When the MASS method is employed, the material can be prepared in one step of 22 min. Despite the short synthesis time, the materials are highly crystalline and have large storage capacity compared to the materials prepared by the conventional solid-state method (Figure 10). Most importantly, complete reduction of Tb4O7 to Tb3+ was achieved by the reducing atmosphere generated in-situ during the synthesis.

3.2. Aluminates

Aluminates are commonly used as a host for LED and energy storage phosphors. The benchmark persistent luminescence material is a doped strontium aluminate with the general formula SrAl2O4:Eu2+, Dy3+ [103] that has the best efficiency to date. Interestingly, the canonical material for pc-LEDs is also an aluminate, Y3Al5O12:Ce3+ [104]. In 2017, Finley et al. [105] used microwave-assisted heating to synthesize SrAl2O4:Eu2+, Dy3+ phosphors starting from reverse micelle precursor. The reverse micelle was prepared in an aqueous metal nitrates solution with a combination of components for emulsion (CTAB, n-Heptane, and 1-butanol). The prepared reverse micelle was submitted to a hybrid microwave heating using activated carbon as a susceptor. The total irradiation time was 14 min, yielding nearly phase-pure materials. The combined reverse micelle with microwave-assisted heating produced particle sizes 70% smaller (d0.5 = 4.2 μm) than the conventional MASS (d0.5 = 14.3 μm). Despite the similar photoluminescence properties, the reverse micelle material has a higher temperature quenching and slightly larger storage capacity.
In 2012, Birkel et al. [83] used the MASS method to prepare RE3Al5O12:Ce3+ (RE: Y, and Lu) starting from RE2O3 and Al2O3, and small quantities of fluoride fluxes. The synthesis is efficient for obtaining highly crystalline garnet phosphors with similar quantum yields to conventional solid-state samples. Usually, luminescent rare earth aluminum garnets need thermal treatment at high temperatures (1500 °C) for several hours to form the material. On the other hand, the MASS synthesis provided an energy saving of 99% with nearly identical optical and structural quality (Figure 11). The quantum yields for the MASS-obtained aluminum garnets were 88% and particle sizes were slightly smaller.

3.3. Silicates

Silicate is an essential class of inorganic host matrices for luminescent materials. The most famous example is the efficient blue-emitting Sr2MgSi2O7:Eu2+ phosphor.
In 2020, Carvalho et al. [57] studied the MASS synthesis of Sr2MgSi2O7:Eu2+, Dy3+ persistent phosphors and microwave irradiation’s effects on the homogeneity and storage capacity. Comparing the conventional and microwave-assisted methods on obtaining the Sr2MgSi2O7:Eu2+, Dy3+ shows that the MASS method provides higher dopant homogeneity in the crystal due to increased ionic diffusion induced by microwave irradiation. However, the local atmosphere generated by carbon susceptor is sometimes inefficient to reduce the activator ions. In this case, decoupled, two-step synthesis is beneficial to obtain Sr2MgSi2O7:Eu2+, Dy3+ materials with high quantum yields, and large storage capacities. In the first step, highly homogeneous materials are obtained due to the increased diffusion in MASS synthesis (Figure 12). The second step reduces the activator ions in mild temperatures achieving optimum Eu2+ concentrations for high-brightness and a large storage capacity.
The MASS method can also be beneficial when the synthesis’s precursor is volatile or decomposes at high temperatures. The swift nature of the dielectric heating allows the reaction to occur before evaporation or degradation takes place. Brgoch et al. [106] obtained Ce3+-doped silicates with complex composition, Na3YSi2O7, using a MASS methodology. The rapid MASS method allows for synthesizing the materials using stoichiometric amounts of the precursors, unlike conventional preparations that must account for sodium carbonate’s evaporation.

3.4. Aluminosilicate

Aluminosilicate consists in AlO4 and SiO4 units interconnected in a 3D framework. Famous example are the zeolites and minerals derived from the sodalite structure. Hackmanite, a sulfide containing sodalite, is a material that exhibits persistent luminescence when doped with Ti ions [107,108]. The white emission has long durations and is the record-long for non-rare earth persistent phosphors [109,110]. The complex structure is usually synthesized using a two-step solid-state synthesis that takes ca. 50 h combined. In 2019, the series of sodalite (Na,M)8Al6Si6O24(Cl,S)2 (M: Li, K, Rb) materials were synthesized using a zeolite precursor applying the MASS method [111]. The structural similarities of the precursors and products, together with the enhanced ionic diffusion induced by MW radiation, reduce the synthesis’ time to 18-20 min. The formation of the material goes through a crystalline intermediate, NaAlSiO4, showing a sharp transition from zeolite to sodalite after 8 min of MW irradiation.
Brgoch et al. studied the MASS synthesis of the Eu-doped (Ba1-xSrx)Al2Si2O8 aluminosilicate materials. The method yields nearly phase-pure materials, which are often difficult to achieve using conventional solid-state methods. Both of the phases, hexagonal and monoclinic, can be synthesized with this method by varying the irradiation time and power levels. Short irradiation (2 min) is enough to synthesize the hexagonal phase, whereas longer times (100 min) lead to the monoclinic phase (Figure 13). The order of the crystallization was interpreted using the Ostwald law wherein the less stable polymorph crystallizes first.

3.5. Sulfides and Oxysulfides

Sulfides and oxysulfides inorganic materials are commonly used as host matrices for luminescent materials. The presence of sulfur, which is bigger and more polarizable compared to oxygen, makes these hosts interesting for emission color tuning of crystal field-sensitive activators, such as Eu2+ and Ce3+.
The synthesis of alkaline earth sulfides is exceedingly complex, usually needing special synthesis setups. For instance, a reducing atmosphere composed by toxic H2S gas is often needed. The usage of poisonous gas limits the synthesis to specialized furnaces. Moreover, the synthesis of the sulfides by conventional method must consider the evaporation of the sulfur during the process. In 2020, Santos et al. [73] synthesized SrS materials doped with various rare-earth ions to compare the co-dopant effect in PeL storage capacity. The MASS synthesis was applied successfully to obtain SrS materials in a single step without using H2S to reduce the sulfur to sulfide anions. Although there were small SrSO4 impurities, crystalline rare-earth-doped SrS materials were obtained with the facile method in 22 min. The decay time of the persistent luminescence was measured in luminance mode, indicating overall good performance of the PeL phosphors.
Sulfur-containing materials are prone to oxidation depending on the synthesis conditions. For instance, La2O2S materials suffer oxidation to La2OSO4 at high temperatures in an oxygen-containing atmosphere [113]. Thus, the synthesis of sulfur-containing materials using MASS synthesis must be finely controlled to avoid excessive temperature rise. In 2018, Carvalho et al. [76] used the MASS method to obtain the RE2O2S (RE: La, Gd, and Y) materials doped with Ti3+/4+ and Mg2+ ions. After the 25 min irradiation, the crystalline phase is mainly the RE2O2S hexagonal phase. The analysis of the sulfur K-edge XANES and FTIR spectra showed rapid oxidation of the elemental sulfur (S0) to sulfate (S6+) in the first 2 min. It is shown that the reduction takes place from 2 to 10 min, where only S2- species are observed (Figure 14). The increased irradiation time promotes partial oxidation of S2- to SO42- species. The results showed a great versatility of the MASS synthesis of sulfur-containing materials where different irradiation times allied with the rapid heating and cooling rates can lead to different sulfur valences.

3.6. Oxometalates

Metalates are the name given to complex anions composed of a metal bonded to a different type of anions. In the case of the oxometallate, a transition metal is bonded to oxygen atoms to form complex anions in various forms, e.g., MnO4, VO43−, CrO42−, WO42−, and MoO42−.
Rare earth-doped tungstate and molybdate matrices are extensively studied as luminescent materials, and are usually obtained by solution-phase methods, such as co-precipitation, sol-gel, hydrothermal, and Pechini. These methods generally require more than one step to separate the material from the solution or heat treating the materials to achieve phase-pure phosphors. Several scheelites have been prepared by using the microwave-assisted solution-phase methods, such as NaLa(MoO4)2 [114], Na(Ca,Sr)(La,Gd)(MoO4)3 [115,116], and (Sr,Pb)(La,Gd,Y)2(MoO4)4 [117]. Solid-state synthesis using microwave heating of oxometallates is scarce, with BaMoO4 [118] and CdWO4 [119] being some examples.
In 2018, Zhai et al. [120] prepared the CaMoO4:Eu3+, Dy3+ scheelite using the MASS method with carbon as susceptor and metal carbonates as precursors. The irradiation program was 30 min at 560 W power level, using a DMO. The synthesized materials possess interesting morphology where micrometric cubes are formed by close-packed quasi-spherical grains.
In 2020, Perera et al. [121] described the MASS synthesis of a series of NaRE(MO4)2 materials (RE: La, Pr, Eu, Dy; M: Mo, W). The synthesis was conducted using pure oxide precursors using various irradiation programs. The total time of the synthesis varied in the interval of 6 to 27 min. The authors reported that the synthesis was robust, fast, and that it was simple to obtain structurally complex metallates. As for the luminescence properties, the materials obtained by MASS synthesis were compared to those from the conventional heating method, and the light output was nearly identical (Figure 15), apart from NaLa0.95Eu0.05(MoO4)2 material that exhibits higher emission intensity for MASS-obtained material.
In 2015, Duée et al. used the microwave-assisted hydrothermal method to obtain luminescent nanovanadates with the formula Y1-xEuxVO4. The method yields small nanoparticles (~10–15 nm) in only 15 min of microwave irradiation. The materials were reported as having efficiencies higher than the state-of-the-art inorganic optical material in liquid phase detection of H2O2.

3.7. Other Materials

Du et al. [122] used the MASS method to obtain a novel Mn4+-doped double-perovskite germanate structure with the formula La2MgGeO6. The approach consisted of two steps, first irradiating the sample with microwaves for 40 min, and then conventional thermal treatment for 3 h. Mn4+ phosphors are emerging as a new class of efficient NIR emitters for bioimaging and night-vision surveillance. However, these materials’ preparation can be complex as the manganese ions can assume different valence states. This approach was used because the hybrid heating using charcoal generates CO gas, which acts as a reductant of Mn4+ to Mn2+, at least partly. The reduction is undesirable because it changes the optical performance and even kills the emission ultimately. The second step of mild temperature heating allows the oxidation of the Mn ions to the tetravalent form. Combining these two effects, it is possible to obtain a highly dispersed dopant, homogeneous crystal structure, and optimized luminescence performance.
Several classes of materials were prepared by microwave-assisted methodologies, and the selected examples are presented in Table 4. The description of the methods, microwave program, and susceptor can be used as a guide to plan a synthetic route using microwave-assisted methodologies.

4. Conclusions and Outlook

In conclusion, the future of luminescent materials obtained by microwave methodologies is bright yet challenging. The microwave-assisted synthesis methodologies are incredibly versatile and are proved to reduce energy consumption and considerably lower the synthesis time of inorganic luminescent materials. Not only is this method more cost-effective than its conventional heating counterpart, but it also has an additional effect on the synthesis, leading to increased reaction rates due to the enhanced ionic diffusion promoted by the electromagnetic field of the MW. Dielectric heating is also an excellent strategy to obtain homogeneous luminescent materials, ideal for illumination purposes, once the intense local heat generation leads to a slight temperature gradient. The rapid heating and quenching of microwave-assisted heating can induce formation of structural defects that allow for high storage capacity of the persistent phosphors.
The shortcoming of the microwave methodologies is sometimes the lack of quantitative comparison between the microwave-obtained materials and other methods, making the advantages challenging to interpret. Furthermore, the methodologies, especially the solid-state one, need to be consolidated in the luminescence community as a recognized method for obtaining phosphor materials. If the radiation hazards are appropriately controlled, novel customized setups can be designed for other methods, such as reflux synthesis using organic solvents and ionic liquids, that might allow obtaining nanoparticles with controlled morphology, such as NaYF4. The design of a dedicated microwave irradiation reactor cavity could also open new avenues for luminescent materials, leading to possibilities of temperature and atmosphere control.
The combination of these properties makes microwave-assisted synthesis an excellent method to obtain luminescent materials that benefit from high crystallinity, well-dispersed dopants, and homogeneous structures. These properties translate to high quantum yields, high brightness, and large storage capacity. Moreover, the microwave-assisted methods allow for fast screening for compositional tailoring, a widespread strategy when tuning the color emission of crystal field-sensitive ions, such as Eu2+ and Ce3+.

Author Contributions

Conceptualization, J.M.d.C., C.C.S.P., M.S.d.N.S., M.C.F.C.F., H.F.d.B.; writing—original draft preparation, J.M.d.C.; writing—review and editing, J.M.d.C., C.C.S.P., M.S.d.N.S., M.C.F.C.F., H.F.d.B.; supervision, J.M.d.C.; project administration, J.M.d.C. All authors have read and agreed to the published version of the manuscript.

Funding

Project FAPESP n° 2017/05195-5, and FAPESP n° 2020/10228-2.

Institutional Review Board Statement

Non-applicable.

Informed Consent Statement

Non-applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Blasse, G.; Grabmaier, B.C. Luminescent Materials; Springer: Berlin/Heidelberg, Germany, 1994; ISBN 978-3-540-58019-5. [Google Scholar]
  2. Ronda, C. Rare-Earth Phosphors: Fundamentals and Applications. In Reference Module in Materials Science and Materials Engineering; Elsevier: Amsterdam, The Netherlands, 2017; ISBN 9780128035818. [Google Scholar]
  3. Malta, O.L.; Brito, H.F.; Menezes, J.F.S.; e Silva, F.R.G.; Alves, S.; Farias, F.S.; de Andrade, A.V.M. Spectroscopic properties of a new light-converting device Eu(thenoyltrifluoroacetonate)3 2(dibenzyl sulfoxide). A theoretical analysis based on structural data obtained from a sparkle model. J. Lumin. 1997, 75, 255–268. [Google Scholar] [CrossRef]
  4. Rodrigues, L.C.V. Preparação e desenvolvimento do mecanismo da luminescência persistente de materiais dopados com íons terras raras. Doctoral Thesis, University of São Paulo, Butanta, Brazil, 2012. [Google Scholar]
  5. Verstraete, R.; Sijbom, H.F.; Joos, J.J.; Korthout, K.; Poelman, D.; Detavernier, C.; Smet, P.F. Red Mn4+-Doped Fluoride Phosphors: Why Purity Matters. ACS Appl. Mater. Interfaces 2018, 10, 18845–18856. [Google Scholar] [CrossRef]
  6. da Silva, M.N.; Carvalho, J.M.; de Abreu Fantini, M.C.; Chiavacci, L.A.; Bourgaux, C. Nanosized ZnGa2O4:Cr3+ Spinels as Highly Luminescent Materials for Bioimaging. ACS Appl. Nano Mater. 2019, 2, 6918–6927. [Google Scholar] [CrossRef]
  7. Ming, H.; Liu, S.; Liu, L.; Peng, J.; Fu, J.; Du, F.; Ye, X. Highly Regular, Uniform K3ScF6:Mn4+ Phosphors: Facile Synthesis, Microstructures, Photoluminescence Properties, and Application in Light-Emitting Diode Devices. ACS Appl. Mater. Interfaces 2018, 10, 19783–19795. [Google Scholar] [CrossRef] [PubMed]
  8. Aarts, L.; van der Ende, B.; Reid, M.F.; Meijerink, A. Downconversion for solar cells in YF3:Pr3+,Yb3+. Spectrosc. Lett. 2010, 43, 373–381. [Google Scholar] [CrossRef]
  9. Terra, I.A.A.; Borrero-González, L.J.; Carvalho, J.M.; Terrile, M.C.; Felinto, M.C.F.C.; Brito, H.F.; Nunes, L.A.O. Spectroscopic properties and quantum cutting in Tb3+–Yb3+ co-doped ZrO2 nanocrystals. J. Appl. Phys. 2013, 113, 073105. [Google Scholar] [CrossRef]
  10. Vergeer, P.; Vlugt, T.J.H.; Kox, M.H.F.; Den Hertog, M.I.; Van Der Herden, J.P.J.M.; Meijerink, A. Quantum cutting by cooperative energy transfer in YbxY1-xPO4:Tb3+. Phys. Rev. B Condens. Matter Mater. Phys. 2005, 71, 1–11. [Google Scholar] [CrossRef]
  11. Yi, G.S.; Chow, G.M. Synthesis of hexagonal-phase NaYF4:Yb,Er and NaYF4:Yb,Tm nanocrystals with efficient up-conversion fluorescence. Adv. Funct. Mater. 2006, 16, 2324–2329. [Google Scholar] [CrossRef]
  12. Hu, L.; Wang, P.; Zhao, M.; Liu, L.; Zhou, L.; Li, B.; Albaqami, F.H.; El-Toni, A.M.; Li, X.; Xie, Y.; et al. Near-infrared rechargeable “optical battery” implant for irradiation-free photodynamic therapy. Biomaterials 2018, 163, 154–162. [Google Scholar] [CrossRef] [PubMed]
  13. Auzel, F. Upconversion and Anti-Stokes Processes with f and d Ions in Solids. Chem. Rev. 2004, 104, 139–173. [Google Scholar] [CrossRef]
  14. Smits, K.; Grigorjeva, L.; Millers, D.; Sarakovskis, A.; Opalinska, A.; Fidelus, J.D.; Lojkowski, W. Europium doped zirconia luminescence. Opt. Mater. 2010, 32, 827–831. [Google Scholar] [CrossRef]
  15. Yang, C.; Fu, L.M.; Wang, Y.; Zhang, J.P.; Wong, W.T.; Ai, X.C.; Qiao, Y.F.; Zou, B.S.; Gui, L.L. A highly luminescent europium complex showing visible-light-sensitized red emission: Direct observation of the singlet pathway. Angew. Chemie. Int. Ed. 2004, 43, 5010–5013. [Google Scholar] [CrossRef]
  16. Pappalardo, R. Calculated quantum yields for photon-cascade emission (PCE) for Pr3+ and Tm3+ in fluoride hosts. J. Lumin. 1976, 14, 159–193. [Google Scholar] [CrossRef]
  17. Piper, W.W.; DeLuca, J.A.; Ham, F.S. Cascade fluorescent decay in Pr3+-doped fluorides: Achievement of a quantum yield greater than unity for emission of visible light. J. Lumin. 1974, 8, 344–348. [Google Scholar] [CrossRef]
  18. Wegh, R.T.; Donker, H.; Oskam, K.D.; Meijerink, A. Visible quantum cutting in LiGdF4:Eu3+ through downconversion. Science 1999, 283, 663–666. [Google Scholar] [CrossRef] [PubMed]
  19. Wegh, R.; Donker, H.; van Loef, E.V.; Oskam, K.; Meijerink, A. Quantum cutting through downconversion in rare-earth compounds. J. Lumin. 2000, 87–89, 1017–1019. [Google Scholar] [CrossRef]
  20. Sommerdijk, J.L.; Bril, A.; de Jager, A.W. Two photon luminescence with ultraviolet excitation of trivalent praseodymium. J. Lumin. 1974, 8, 341–343. [Google Scholar] [CrossRef]
  21. Bloembergen, N. Solid State Infrared Quantum Counters. Phys. Rev. Lett. 1959, 2, 84–85. [Google Scholar] [CrossRef]
  22. Zhou, B.; Shi, B.; Jin, D.; Liu, X. Controlling upconversion nanocrystals for emerging applications. Nat. Nanotechnol. 2015, 10, 924–936. [Google Scholar] [CrossRef]
  23. Xu, J.; Tanabe, S. Persistent luminescence instead of phosphorescence: History, mechanism, and perspective. J. Lumin. 2019, 205, 581–620. [Google Scholar] [CrossRef]
  24. Van den Eeckhout, K.; Poelman, D.; Smet, P.F. Persistent luminescence in Non-Eu2+-Doped compounds: A review. Materials 2013, 6, 2789. [Google Scholar] [CrossRef] [Green Version]
  25. Van den Eeckhout, K.; Smet, P.F.; Poelman, D. Persistent Luminescence in Eu2+-Doped Compounds: A Review. Materials 2010, 3, 2536–2566. [Google Scholar] [CrossRef] [Green Version]
  26. Im, W.B.; George, N.; Kurzman, J.; Brinkley, S.; Mikhailovsky, A.; Hu, J.; Chmelka, B.F.; Denbaars, S.P.; Seshadri, R. Efficient and color-tunable oxyfluoride solid solution phosphors for solid-state white lighting. Adv. Mater. 2011, 23, 2300–2305. [Google Scholar] [CrossRef] [PubMed]
  27. Im, W.B.; Brinkley, S.; Hu, J.; Mikhailovsky, A.; Denbaars, S.P.; Seshadri, R. Sr2.975-xBaxCe0.025AlO4F: A highly efficient green-emitting oxyfluoride phosphor for solid state white lighting. Chem. Mater. 2010, 22, 2842–2849. [Google Scholar] [CrossRef]
  28. Dorenbos, P. Ce3+ 5d-centroid shift and vacuum referred 4f-electron binding energies of all lanthanide impurities in 150 different compounds. J. Lumin. 2013, 135, 93–104. [Google Scholar] [CrossRef]
  29. Yamaga, M.; Yosida, T.; Hara, S.; Kodama, N.; Henderson, B. Optical and electron spin resonance spectroscopy of Ti3+ and Ti4+ in Al2O3. J. Appl. Phys. 1994, 75, 1111–1117. [Google Scholar] [CrossRef]
  30. Katumo, N.; Gao, G.; Laufer, F.; Richards, B.S.; Howard, I.A. Smartphone-Based Luminescent Thermometry via Temperature-Sensitive Delayed Fluorescence from Gd2 O2 S:Eu3+. Adv. Opt. Mater. 2020, 8, 2000507. [Google Scholar] [CrossRef]
  31. Chen, R.; Pagonis, V. Modelling thermal activation characteristics of the sensitization of thermoluminescence in quartz. J. Phys. D Appl. Phys. 2004, 37, 159–164. [Google Scholar] [CrossRef] [Green Version]
  32. Chen, R.; Pagonis, V.; Lawless, J.L. Evaluated thermoluminescence trapping parameters-What do they really mean? Radiat. Meas. 2016, 91, 21–27. [Google Scholar] [CrossRef]
  33. Kristianpoller, N.; Chen, R.; Israeli, M. Dose dependence of thermoluminescence peaks. J. Phys. D Appl. Phys. 1974, 7, 1063–1072. [Google Scholar] [CrossRef]
  34. Dong, H.; Du, S.R.; Zheng, X.Y.; Lyu, G.M.; Sun, L.D.; Li, L.D.; Zhang, P.Z.; Zhang, C.; Yan, C.H. Lanthanide Nanoparticles: From Design toward Bioimaging and Therapy. Chem. Rev. 2015, 115, 10725–10815. [Google Scholar] [CrossRef]
  35. Wang, Q.; Liu, N.; Hou, Z.; Shi, J.; Su, X.; Sun, X. Radioiodinated Persistent Luminescence Nanoplatform for Radiation-Induced Photodynamic Therapy and Radiotherapy. Adv. Healthc. Mater. 2021, 10, 2000802. [Google Scholar] [CrossRef] [PubMed]
  36. Liu, J.; Lécuyer, T.; Seguin, J.; Mignet, N.; Scherman, D.; Viana, B.; Richard, C. Imaging and therapeutic applications of persistent luminescence nanomaterials. Adv. Drug Deliv. Rev. 2019, 138, 193–210. [Google Scholar] [CrossRef]
  37. Tan, H.; Wang, T.; Shao, Y.; Yu, C.; Hu, L. Crucial Breakthrough of Functional Persistent Luminescence Materials for Biomedical and Information Technological Applications. Front. Chem. 2019, 7. [Google Scholar] [CrossRef]
  38. Smet, P.F.; Parmentier, A.B.; Poelman, D. Selecting Conversion Phosphors for White Light-Emitting Diodes. J. Electrochem. Soc. 2011, 158, R37. [Google Scholar] [CrossRef] [Green Version]
  39. Schubert, E.F.; Kim, J.K. Solid-state light sources getting smart. Science 2005, 308, 1274–1278. [Google Scholar] [CrossRef]
  40. George, N.C.; Denault, K.A.; Seshadri, R. Phosphors for Solid-State White Lighting. Annu. Rev. Mater. Res. 2013, 43, 481–501. [Google Scholar] [CrossRef]
  41. Sijbom, H.F.; Verstraete, R.; Joos, J.J.; Poelman, D.; Smet, P.F. K2SiF6:Mn4+ as a red phosphor for displays and warm-white LEDs: A review of properties and perspectives. Opt. Mater. Express 2017, 7, 3332. [Google Scholar] [CrossRef]
  42. Saradhi, M.P.; Boudin, S.; Varadaraju, U.V.; Raveau, B. A new BaB2Si2O8:Eu2/Eu3, Tb3phosphor Synthesis and photoluminescence properties. J. Solid State Chem. 2010, 183, 2496–2500. [Google Scholar] [CrossRef]
  43. Hızal Abacı, Ö.C.; Mete, E.; Esenturk, O.; Yılmaz, A. Tunable optical properties and DFT calculations of RE3+ codoped LaBO3 phosphors. Opt. Mater. 2019, 98, 3–12. [Google Scholar] [CrossRef]
  44. Andrade, A.B.; Bispo, G.F.C.; Macedo, Z.S.; Valerio, M.E.G. Synthesis and characterization of luminescent Ln3+ (Ln = Eu, Tb and Dy)-doped LiYF4 microcrystals produced by a facile microwave-assisted hydrothermal method. J. Lumin. 2020, 219, 116843. [Google Scholar] [CrossRef]
  45. Chuang, Y.J.; Zhen, Z.; Zhang, F.; Liu, F.; Mishra, J.P.; Tang, W.; Chen, H.; Huang, X.; Wang, L.; Chen, X.; et al. Photostimulable near-infrared persistent luminescent nanoprobes for ultrasensitive and longitudinal deep-tissue bio-imaging. Theranostics 2014, 4, 1112–1122. [Google Scholar] [CrossRef] [Green Version]
  46. Kang, F.; Sun, G.; Boutinaud, P.; Wu, H.; Ma, F.-X.; Lu, J.; Gan, J.; Bian, H.; Gao, F.; Xiao, S. Recent advances and prospects of persistent luminescent materials as inner secondary self-luminous light source for photocatalytic applications. Chem. Eng. J. 2021, 403, 126099. [Google Scholar] [CrossRef]
  47. Van der Heggen, D.; Joos, J.J.; Rodríguez Burbano, D.C.; Capobianco, J.A.; Smet, P.F. Counting the photons: Determining the absolute storage capacity of persistent phosphors. Materials 2017, 10, 867. [Google Scholar] [CrossRef]
  48. Van der Heggen, D.; Joos, J.; Smet, P.F. On the importance of evaluating the intensity dependency of the quantum efficiency: Impact on LEDs and persistent phosphors. ACS Photonics 2018, 5, 4529–4537. [Google Scholar] [CrossRef]
  49. Bhattacharya, M.; Basak, T. A review on the susceptor assisted microwave processing of materials. Energy 2016, 97, 306–338. [Google Scholar] [CrossRef]
  50. Schiffmann, R.F. Microwave and dielectric drying. In Handbook of Industrial Drying, 4th ed.; CRC Press: Boca Raton, FL, USA, 2014; ISBN 9781466596665. [Google Scholar]
  51. Kitchen, H.J.; Vallance, S.R.; Kennedy, J.L.; Tapia-Ruiz, N.; Carassiti, L.; Harrison, A.; Whittaker, A.G.; Drysdale, T.D.; Kingman, S.W.; Gregory, D.H. Modern microwave methods in solid-state inorganic materials chemistry: From fundamentals to manufacturing. Chem. Rev. 2014, 114, 1170–1206. [Google Scholar] [CrossRef]
  52. Rao, K.J.; Ramesh, P.D. Use of microwaves for the synthesis and processing of materials. Bull. Mater. Sci. 1995, 18, 447–465. [Google Scholar] [CrossRef]
  53. Naitoh, K.; Takizawa, T.; Matsuse, T. Controlled microwave irradiation for the synthesis of YBa2Cu3O7-x superconductors. Jpn. J. Appl. Phys. Part 2 Lett. 1999. [Google Scholar] [CrossRef]
  54. Vaidhyanathan, B.; Balaji, K.; Rao, K.J. Microwave-Assisted Solid-State Synthesis of Oxide Ion Conducting Stabilized Bismuth Vanadate Phases. Chem. Mater. 1998, 10, 3400–3404. [Google Scholar] [CrossRef]
  55. Rao, K.J.; Vaidhyanathan, B.; Ganguli, M.; Ramakrishnan, P.A. Synthesis of inorganic solids using microwaves. Chem. Mater. 1999, 11, 882–895. [Google Scholar] [CrossRef]
  56. Kappe, C.O. Controlled microwave heating in modern organic synthesis. Angew. Chem. Int. Ed. 2004, 43, 6250–6284. [Google Scholar] [CrossRef]
  57. Miranda De Carvalho, J.; Van Der Heggen, D.; Martin, L.I.D.J.; Smet, P.F. Microwave-assisted synthesis followed by a reduction step: Making persistent phosphors with a large storage capacity. Dalt. Trans. 2020, 49, 4518–4527. [Google Scholar] [CrossRef] [PubMed]
  58. Sebastian, M.T. Dielectric Materials for Wireless Communication; Elsevier: Amsterdam, The Netherlands, 2008; ISBN 9780080453309. [Google Scholar]
  59. Hotta, M.; Hayashi, M.; Lanagan, M.T.; Agrawal, D.K.; Nagata, K. Complex permittivity of graphite, carbon black and coal powders in the ranges of X-band frequencies (8.2 to 12.4 ghz) and between 1 and 10 ghz. ISIJ Int. 2011, 51, 1766–1772. [Google Scholar] [CrossRef] [Green Version]
  60. Balanis, C.A. (Ed.) Advanced Engineering Electromagnetics; Wiley: Hoboken, NJ, USA, 2007; ISBN 9780470589489. [Google Scholar]
  61. Gupta, M.; Eugene, W.W.L. Microwaves and Metals; John Wiley & Sons: Hoboken, NJ, USA, 2011; ISBN 9780470822722. [Google Scholar]
  62. Pedroso, C.C.S.; Carvalho, J.M.; Rodrigues, L.C.V.; Hölsä, J.; Brito, H.F. Rapid and Energy-Saving Microwave-Assisted Solid-State Synthesis of Pr3+-, Eu3+-, or Tb3+-Doped Lu2O3Persistent Luminescence Materials. ACS Appl. Mater. Interfaces 2016, 8, 19593–19604. [Google Scholar] [CrossRef] [PubMed]
  63. Levin, E.E.; Grebenkemper, J.H.; Pollock, T.M.; Seshadri, R. Protocols for High Temperature Assisted-Microwave Preparation of Inorganic Compounds. Chem. Mater. 2019, 31, 7151–7159. [Google Scholar] [CrossRef] [Green Version]
  64. Bykov, Y.V.; Rybakov, K.I.; Semenov, V.E. High-temperature microwave processing of materials. J. Phys. D Appl. Phys. 2001, 34, R55–R75. [Google Scholar] [CrossRef]
  65. Ku, H.S.; Ball, J.A.R.; Siores, E. Review-Microwave Processing of Materials: Part III. HKIE Trans. 2001, 8, 44–50. [Google Scholar] [CrossRef] [Green Version]
  66. Metaxas, A.C.; Meredith, R.J. Dielectric loss. In Industrial Microwave Heating; IET: Six Hills Way, UK, 2011; pp. 5–25. [Google Scholar]
  67. Li, W.; Adlung, M.; Zhang, Q.; Wickleder, C.; Schmedt auf der Günne, J. A Guide to Brighter Phosphors-Linking Luminescence Properties to Doping Homogeneity Probed by NMR. ChemPhysChem 2019, 20, 3245–3250. [Google Scholar] [CrossRef] [Green Version]
  68. Clark, D.E.; Folz, D.C.; West, J.K. Processing materials with microwave energy. Mater. Sci. Eng. A 2000, 287, 153–158. [Google Scholar] [CrossRef]
  69. Gusarov, V.V. Fast Solid-Phase Chemical Reactions. Russ. J. Gen. Chem. 1997, 67, 1846–1851. [Google Scholar]
  70. Whittaker, A.G. Diffusion in microwave-heated ceramics. Chem. Mater. 2005, 17, 3426–3432. [Google Scholar] [CrossRef]
  71. Vanetsev, A.S.; Baranchikov, A.E.; Tret’yakov, Y.D. Kinetics of microwave-enhanced solid-phase reaction of NiFe2O4 formation. Russ. J. Inorg. Chem. 2008, 53, 495–498. [Google Scholar] [CrossRef]
  72. Rödiger, K.; Dreyer, K.; Gerdes, T.; Willert-Porada, M. Microwave sintering of hardmetals. Int. J. Refract. Met. Hard Mater. 1998, 16, 409–416. [Google Scholar] [CrossRef]
  73. dos Santos, D.O.A.; Giordano, L.; Barbará, M.A.S.G.; Portes, M.C.; Pedroso, C.C.S.; Teixeira, V.C.; Lastusaari, M.; Rodrigues, L.C.V. Abnormal co-doping effect on the red persistent luminescence SrS:Eu2+,RE3+ materials. Dalt. Trans. 2020, 49, 16386–16393. [Google Scholar] [CrossRef]
  74. Chandramouli, V.; Anthonysamy, S.; Vasudeva Rao, P.R.; Divakar, R.; Sundararaman, D. Microwave synthesis of solid solutions of urania and thoria-A comparative study. J. Nucl. Mater. 1998, 254, 55–64. [Google Scholar] [CrossRef]
  75. Ganesh, I.; Johnson, R.; Rao, G.V.N.; Mahajan, Y.R.; Madavendra, S.S.; Reddy, B.M. Microwave-assisted combustion synthesis of nanocrystalline MgAl2O4 spinel powder. Ceram. Int. 2005, 31, 67–74. [Google Scholar] [CrossRef]
  76. Miranda De Carvalho, J.; Pedroso, C.C.S.; Machado, I.P.; Hölsä, J.; Rodrigues, L.C.V.; Głuchowski, P.; Lastusaari, M.; Brito, H.F. Persistent luminescence warm-light LEDs based on Ti-doped RE2O2S materials prepared by rapid and energy-saving microwave-assisted synthesis. J. Mater. Chem. C 2018, 6, 8897–8905. [Google Scholar] [CrossRef]
  77. Machado, I.P.; Teixeira, V.C.; Pedroso, C.C.S.; Brito, H.F.; Rodrigues, L.C.V. X-ray scintillator Gd2O2S:Tb3+ materials obtained by a rapid and cost-effective microwave-assisted solid-state synthesis. J. Alloys Compd. 2019, 777, 638–645. [Google Scholar] [CrossRef]
  78. Lee, W.C.; Liu, K.S.; Lin, I.N. Microwave sintering Pb(Zr0.52Ti0.48)O3 piezoelectric ceramics. Ferroelectrics 2001, 262, 293–298. [Google Scholar] [CrossRef]
  79. Walkiewicz, J.W.; Kazonich, G.; McGill, S.L. Micorwave Heating Characteristics of Selected Minerals and Compounds. Miner. Metall. Process. 1988, 5, 39–42. [Google Scholar] [CrossRef]
  80. Song, B.; Zhao, B.; Fan, L.; Fan, B.; Wang, H.; Guo, X.; Zhang, R. Investigation on heating behavior during the preparation of SiC crystals by microwave sintering. Int. J. Appl. Ceram. Technol. 2017, 14, 880–888. [Google Scholar] [CrossRef]
  81. Rao, K.V.; Smakula, A. Dielectric properties of cobalt oxide, nickel oxide, and their mixed crystals. J. Appl. Phys. 1965, 36, 2031–2038. [Google Scholar] [CrossRef]
  82. Kulesza, D.; Bolek, P.; Bos, A.J.J.; Zych, E. Lu2O3-based storage phosphors. An (in)harmonious family. Coord. Chem. Rev. 2016, 325, 29–40. [Google Scholar] [CrossRef] [Green Version]
  83. Birkel, A.; Denault, K.A.; George, N.C.; Doll, C.E.; Héry, B.; Mikhailovsky, A.A.; Birkel, C.S.; Hong, B.C.; Seshadri, R. Rapid microwave preparation of highly efficient Ce3+-substituted garnet phosphors for solid state white lighting. Chem. Mater. 2012, 24, 1198–1204. [Google Scholar] [CrossRef]
  84. Game, D.N.; Palan, C.B.; Ingale, N.B.; Omanwar, S.K. Microwave-assisted synthesis and photoluminescence properties of ZnS:Pb2+ nanophosphor for solid-state lighting. Bull. Mater. Sci. 2017, 40, 1441–1445. [Google Scholar] [CrossRef] [Green Version]
  85. Kamińska, I.; Jankowski, D.; Sikora, B.; Kowalik, P.; Minikayev, R.; Wojciechowski, T.; Chojnacki, M.; Sobczak, K.; Rybusiński, J.; Szczytko, J.; et al. Structural, optical and magnetic properties of Y3−0.02−xEr 0.02YbxAl5O12 (0 <x <0.20) nanocrystals: Effect of Yb content. Nanotechnology 2020, 31, 225711. [Google Scholar] [CrossRef]
  86. Liu, G.; Li, J.; Chen, K. Combustion synthesis of refractory and hard materials: A review. Int. J. Refract. Met. Hard Mater. 2013, 39, 90–102. [Google Scholar] [CrossRef]
  87. Patil, K.C.; Aruna, S.T.; Ekambaram, S. Combustion synthesis. Curr. Opin. Solid State Mater. Sci. 1997. [Google Scholar] [CrossRef]
  88. Patil, K.C.; Aruna, S.T.; Mimani, T. Combustion synthesis: An update. Curr. Opin. Solid State Mater. Sci. 2002, 6, 507–512. [Google Scholar] [CrossRef]
  89. Varma, A.; Rogachev, A.S.; Mukasyan, A.S.; Hwang, S. Combustion Synthesis of Advanced Materials: Principles and Applications. In Advances in Chemical Engineering; Academic Press: Cambridge, MA, USA, 1998; pp. 79–226. [Google Scholar]
  90. Kamińska, I.; Fronc, K.; Sikora, B.; Mouawad, M.; Siemiarczuk, A.; Szewczyk, M.; Sobczak, K.; Wojciechowski, T.; Zaleszczyk, W.; Minikayev, R.; et al. Upconverting/magnetic: Gd2O3:(Er3+,Yb3+,Zn2+) nanoparticles for biological applications: Effect of Zn2+ doping. RSC Adv. 2015, 5, 78361–78373. [Google Scholar] [CrossRef]
  91. Meng, L.-Y.; Wang, B.; Ma, M.-G.; Lin, K.-L. The progress of microwave-assisted hydrothermal method in the synthesis of functional nanomaterials. Mater. Today Chem. 2016, 1–2, 63–83. [Google Scholar] [CrossRef]
  92. Johnson, G.M.; Mead, P.J.; Weller, M.T. Synthesis of a range of anion-containing gallium and germanium sodalites. Microporous Mesoporous Mater. 2000, 38, 445–460. [Google Scholar] [CrossRef]
  93. Misch, L.M.; Birkel, A.; Figg, C.A.; Fors, B.P.; Hawker, C.J.; Stucky, G.D.; Seshadri, R. Rapid microwave-assisted sol–gel preparation of Pd-substituted LnFeO3 (Ln = Y, La): Phase formation and catalytic activity. Dalt. Trans. 2014, 43, 2079–2087. [Google Scholar] [CrossRef]
  94. Teston, E.; Richard, S.; Maldiney, T.; Lièvre, N.; Wang, G.Y.; Motte, L.; Richard, C.; Lalatonne, Y. Non-aqueous sol-gel synthesis of ultra small persistent luminescence nanoparticles for near-infrared in vivo imaging. Chem. A Eur. J. 2015, 21, 7350–7354. [Google Scholar] [CrossRef]
  95. Bilecka, I.; Djerdj, I.; Niederberger, M. One-minute synthesis of crystalline binary and ternary metal oxide nanoparticles. Chem. Commun. 2008, 7, 886–888. [Google Scholar] [CrossRef] [PubMed]
  96. Bi, J.; Sun, L.; Wei, Q.; Zhang, K.; Zhu, L.; Wei, S.; Liao, D.; Sun, J. Rapid ultrasonic-microwave assisted synthesis of Eu3+ doped Y2O3 nanophosphors with enhanced luminescence properties. J. Mater. Res. Technol. 2020, 9, 9523–9530. [Google Scholar] [CrossRef]
  97. Thongtem, T.; Phuruangrat, A.; Thongtem, S. Microwave-assisted synthesis and characterization of SrMoO4 and SrWO4 nanocrystals. J. Nanoparticle. Res. 2010, 12, 2287–2294. [Google Scholar] [CrossRef]
  98. Birkel, A.; DeCino, N.A.; George, N.C.; Hazelton, K.A.; Hong, B.-C.; Seshadri, R. Eu2+-doped M2SiO4 (M = Ca, Ba) phosphors prepared by a rapid microwave-assisted sol–gel method: Phase formation and optical properties. Solid State Sci. 2013, 19, 51–57. [Google Scholar] [CrossRef]
  99. Duée, N.; Ambard, C.; Pereira, F.; Portehault, D.; Viana, B.; Vallé, K.; Autissier, D.; Sanchez, C. New Synthesis Strategies for Luminescent YVO4:Eu and EuVO4 Nanoparticles with H2O2 Selective Sensing Properties. Chem. Mater. 2015, 27, 5198–5205. [Google Scholar] [CrossRef]
  100. Gonçalves, R.F.; Longo, E.; Marques, A.P.A.; Silva, M.D.P.; Cavalcante, L.S.; Nogueira, I.C.; Pinatti, I.M.; Pereira, P.F.S.; Godinho, M.J. Structural investigation and photoluminescent properties of ZnWO4:Dy3+ nanocrystals. J. Mater. Sci. Mater. Electron. 2017, 28, 15466–15479. [Google Scholar] [CrossRef]
  101. Zych, E.; Dereń, P.; Strȩk, W.; Meijerink, A.; Mielcarek, W.; Domagala, K. Preparation, X-ray analysis and spectroscopic investigation of nanostructured Lu2O3:Tb. J. Alloys Compd. 2001, 323–324, 8–12. [Google Scholar] [CrossRef]
  102. Trojan-Piegza, J.; Zych, E. Afterglow Luminescence of Lu2O3:Eu Ceramics Synthesized at Different Atmospheres. J. Phys. Chem. C 2010, 114, 4215–4220. [Google Scholar] [CrossRef]
  103. Matsuzawa, T.; Aoki, Y.; Takeuchi, N.; Murayama, Y. A New Long Phosphorescent Phosphor with High Brightness, SrAl2O4:Eu2+,Dy3+. J. Electrochem. Soc. 1996, 143, 2670–2673. [Google Scholar] [CrossRef]
  104. Blasse, G.; Bril, A. A new phosphor for flying-spot cathode-ray tubes for color television: Yellow-emittng Y3Al5O12–Ce3+. Appl. Phys. Lett. 1967, 11, 53–55. [Google Scholar] [CrossRef]
  105. Finley, E.; Paterson, A.S.; Cobb, A.; Willson, R.C.; Brgoch, J. Reducing particle size of persistent luminescent SrAl2O4:Eu2+,Dy3+ via microwave-assisted, reverse micelle synthesis. Opt. Mater. Express 2017, 7, 2597. [Google Scholar] [CrossRef]
  106. Brgoch, J.; Borg, C.K.H.; Denault, K.A.; Douglas, J.R.; Amanda Strom, T.; DenBaars, S.P.; Seshadri, R. Rapid microwave preparation of cerium-substituted sodium yttrium silicate phosphors for solid state white lighting. Solid State Sci. 2013, 26, 115–120. [Google Scholar] [CrossRef]
  107. Norrbo, I.; Gluchowski, P.; Hyppänen, I.; Laihinen, T.; Laukkanen, P.; Mäkelä, J.; Mamedov, F.; Santos, H.S.; Sinkkonen, J.; Tuomisto, M.; et al. Mechanisms of Tenebrescence and Persistent Luminescence in Synthetic Hackmanite Na8Al6Si6O24(Cl,S)2. ACS Appl. Mater. Interfaces 2016, 8, 11592–11602. [Google Scholar] [CrossRef] [PubMed]
  108. Agamah, C.; Vuori, S.; Colinet, P.; Norrbo, I.; Miranda de Carvalho, J.; Okada Nakamura, L.; Lindblom, J.; van Goethem, L.; Emmermann, A.; Saarinen, T.; et al. Hackmanite—The Natural Glow-in-the-Dark Material. Chem. Mater. 2020, 32, 8895–8905. [Google Scholar] [CrossRef]
  109. Norrbo, I.; Carvalho, J.M.; Laukkanen, P.; Mäkelä, J.; Mamedov, F.; Peurla, M.; Helminen, H.; Pihlasalo, S.; Härmä, H.; Sinkkonen, J. Lanthanide and Heavy Metal Free Long White Persistent Luminescence from Ti Doped Li–Hackmanite: A Versatile, Low-Cost Material. Adv. Funct. Mater. 2017, 27, 1696547. [Google Scholar] [CrossRef]
  110. Norrbo, I.; Gluchowski, P.; Paturi, P.; Sinkkonen, J.; Lastusaari, M. Persistent Luminescence of Tenebrescent Na8Al6Si6O24(Cl,S)2: Multifunctional Optical Markers. Inorg. Chem. 2015, 54, 7717–7724. [Google Scholar] [CrossRef] [PubMed]
  111. Carvalho, J.M.; Norrbo, I.; Ando, R.A.; Brito, H.F.; Fantini, M.C.A.; Lastusaari, M. Fast, low-cost preparation of hackmanite minerals with reversible photochromic behavior using a microwave-assisted structure-conversion method. Chem. Commun. 2018, 54, 7326–7329. [Google Scholar] [CrossRef]
  112. Brgoch, J.; Kloß, S.D.; Denault, K.A.; Seshadri, R. Accessing (Ba1-x Srx)Al2Si2O8:Eu Phosphors for Solid State White Lighting via Microwave-assisted Preparation: Tuning Emission Color by Coordination Environment. Zeitschrift für Anorg. Allg. Chem. 2014, 640, 1182–1189. [Google Scholar] [CrossRef]
  113. Machida, M.; Kawano, T.; Eto, M.; Zhang, D.; Ikeue, K. Ln dependence of the large-capacity oxygen storage/release property of Ln oxysulfate/oxysulfide systems. Chem. Mater. 2007, 19, 954–960. [Google Scholar] [CrossRef]
  114. Zhang, J.; Wang, X.; Zhang, X.; Zhao, X.; Liu, X.; Peng, L. Microwave synthesis of NaLa(MoO4)2 microcrystals and their near-infrared luminescent properties with lanthanide ion doping (Er3+, Nd3+, Yb3+). Inorg. Chem. Commun. 2011, 14, 1723–1727. [Google Scholar] [CrossRef]
  115. Lim, C.S.; Aleksandrovsky, A.S.; Molokeev, M.S.; Oreshonkov, A.S.; Ikonnikov, D.A.; Atuchin, V.V. Triple molybdate scheelite-type upconversion phosphor NaCaLa(MoO4)3:Er3+/Yb3+: Structural and spectroscopic properties. Dalt. Trans. 2016, 45, 15541–15551. [Google Scholar] [CrossRef] [Green Version]
  116. Lim, C.S.; Aleksandrovsky, A.S.; Molokeev, M.S.; Oreshonkov, A.S.; Atuchin, V.V. Microwave synthesis and spectroscopic properties of ternary scheelite-type molybdate phosphors NaSrLa(MoO4)3:Er3+,Yb3+. J. Alloys Compd. 2017, 713, 156–163. [Google Scholar] [CrossRef]
  117. Lim, C.S.; Aleksandrovsky, A.S.; Atuchin, V.V.; Molokeev, M.S.; Oreshonkov, A.S. Microwave sol-gel synthesis, microstructural and spectroscopic properties of scheelite-type ternary molybdate upconversion phosphor NaPbLa(MoO4)3:Er3+/Yb3+. J. Alloys Compd. 2020, 826, 152095. [Google Scholar] [CrossRef] [Green Version]
  118. Klinbumrung, A.; Phuruangrat, A.; Thongtem, T.; Thongtem, S. Synthesis, Characterization and Optical Properties of BaMoO4 Synthesized by Microwave Induced Plasma Method. Russ. J. Inorg. Chem. 2018, 63, 725–731. [Google Scholar] [CrossRef]
  119. Sofronov, D.; Sofronova, E.; Starikov, V.; Baymer, V.; Kudin, K.; Matejchenko, P.; Mamalis, A.; Lavrynenko, S. Microwave Synthesis of Tetragonal Phase CdWO4. Mater. Manuf. Process. 2012, 27, 490–493. [Google Scholar] [CrossRef]
  120. Zhai, Y.; Zhao, X.; Liu, C.; Song, P.; Jing, X.; Han, Y.; Wang, J. CaMoO4:Dy3+,Eu3+ phosphors: Microwave synthesis, characterization, tunable luminescence properties and energy transfer mechanism. Optik 2018, 164, 433–442. [Google Scholar] [CrossRef]
  121. Perera, S.S.; Munasinghe, H.N.; Yatooma, E.N.; Rabuffetti, F.A. Microwave-assisted solid-state synthesis of NaRE(MO4)2 phosphors (RE = La, Pr, Eu, Dy; M = Mo, W). Dalt. Trans. 2020, 49, 7914–7919. [Google Scholar] [CrossRef] [PubMed]
  122. Du, J.; Poelman, D. Facile synthesis of mn4+-activated double perovskite germanate phosphors with near-infrared persistent luminescence. Nanomaterials 2019, 9, 1759. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Chen, Z.; Yan, Y.; Liu, J.; Yin, Y.; Wen, H.; Zao, J.; Liu, D.; Tian, H.; Zhang, C.; Li, S. Microwave induced solution combustion synthesis of nano-sized phosphors. J. Alloys Compd. 2009, 473, 13–16. [Google Scholar] [CrossRef]
  124. Barros, B.S.; De Oliveira, R.S.; Kulesza, J.; Melo, V.R.M.; Melo, D.M.A.; Alves, S. Reddish-orange Ca3-xAl2O6:xEu3+ nanophosphors: Fast synthesis and photophysical properties. J. Phys. Chem. Solids 2015, 78, 90–94. [Google Scholar] [CrossRef]
  125. Zhang, P.; Lingxia, L.; Yuming, T. The low temperature synthesis of Eu2+ and Dy3+ activated Sr3Al2O6 nanophosphors by microwave method. Phys. B Condens. Matter 2009, 404, 4286–4289. [Google Scholar] [CrossRef]
  126. Lau, K.S.; Hassan, Z.; Lim, W.F.; Mohammad, S.M.; Quah, H.J. Effect of microwave time on the structural and luminescence properties of YAG:Ce prepared by microwave solution combustion (MSC) synthesis. Optik 2020, 212, 164437. [Google Scholar] [CrossRef]
  127. Anand, G.T.; Kennedy, L.J. One-pot microwave combustion synthesis of porous Zn1-xCuxAl2O4 (0 ≤ x ≤ 0.5) Spinel Nanostructures. J. Nanosci. Nanotechnol. 2013, 13, 3096–3103. [Google Scholar] [CrossRef]
  128. Tong, Y.-Y.; Li, Y.-F.; Chen, R.; Yang, R.; Sun, L.-L.; Zhang, S.; Fu, Y.-Y. Luminescence characteristics and Förster resonance energy transfer of Y and Bi codoped Al6Si2O13:Eu red nanophosphors obtained by an energy-saving microwave-assisted synthesis. J. Lumin. 2020, 224, 117234. [Google Scholar] [CrossRef]
  129. Parhi, P.; Manivannan, V. Novel microwave assisted solid state metathesis synthesis of KMF3 (M = Zn, Mn, Mg, and Co). Mater. Lett. 2008, 62, 3468–3470. [Google Scholar] [CrossRef]
  130. Guzzetta, F.; Roig, A.; Julián-López, B. Ultrafast Synthesis and Coating of High-Quality β-NaYF4:Yb3+,Ln3+ Short Nanorods. J. Phys. Chem. Lett. 2017, 8, 5730–5735. [Google Scholar] [CrossRef] [PubMed]
  131. Makhov, V.N.; Vanetsev, A.S.; Khaidukov, N.M.; Yin, M.; Wei, X.T.; Kotlov, A.; Belsky, A.N. Intrinsic and impurity luminescence of rare earth ions doped KYF4 nanophosphors. Radiat. Meas. 2013, 56, 393–396. [Google Scholar] [CrossRef] [Green Version]
  132. Yang, M.; Deng, G.; Hou, T.; Jia, X.; Wang, Y.; Wang, Q.; Li, B.; Liu, J.; Liu, X. Facile microwave-assisted synthesis of Zn2GeO4:Mn2+,Yb3+ uniform nanorods and near-infrared down-conversion properties. Opt. Mater. 2017, 64, 152–159. [Google Scholar] [CrossRef]
  133. Lim, C.S.; Atuchin, V.V.; Aleksandrovsky, A.S.; Molokeev, M.S.; Oreshonkov, A.S. Incommensurately modulated structure and spectroscopic properties of CaGd2(MoO4)4:Ho3+/Yb3+ phosphors for up-conversion applications. J. Alloys Compd. 2017, 695, 737–746. [Google Scholar] [CrossRef] [Green Version]
  134. Liu, J.; Liu, K.; Gao, H.; Liu, Y.; Yang, C.; Liu, Z. Synthesis and luminescence properties of CaMoO4: Er3+/Yb3+ nanoparticles. J. Mater. Sci. Mater. Electron. 2015, 26, 3380–3383. [Google Scholar] [CrossRef]
  135. Liu, X.; Lv, L.; Huang, S.; Su, Y.; Wang, X. Microwave-assisted synthesis and luminescence properties of Cd1-xEuxMoO4 red phosphor. J. Nanosci. Nanotechnol. 2014, 14, 3618–3622. [Google Scholar] [CrossRef] [PubMed]
  136. Çelik Gül, G.; Kurtuluş, F. Cr and Co doped Cs5Bi(MoO4)4: Microwave assisted synthesis, characterization and specification of optical properties. Optik 2017, 132, 153–163. [Google Scholar] [CrossRef]
  137. Huang, R.; Wang, Q.; Liao, J.; Weixiong, Y.O.U. Microwave hydrothermal synthesis and upconversion luminescence properties of Yb3+/Tm3+ co-doped NaY(MoO4)2 phosphor. Bull. Mater. Sci. 2017, 40, 1447–1453. [Google Scholar] [CrossRef] [Green Version]
  138. Lim, C.S. Microwave sol-gel derived Ho3+/Yb3+ co-doped NaCaGd(MoO4)3 phosphors and their upconversion photoluminescence. J. Korean Ceram. Soc. 2016, 53, 456–462. [Google Scholar] [CrossRef] [Green Version]
  139. Lim, C.S. Synthesis of NaCaLa(MoO4)3:Ho3+/Yb3+ phosphors via microwave Sol-Gel route and their upconversion photoluminescence properties. Korean J. Mater. Res. 2016, 26, 363–369. [Google Scholar] [CrossRef]
  140. Lim, C.S. Microwave sol-gel derived NaCaGd(MoO4)3:Er3+/Yb3+ phosphors and their upconversion photoluminescence properties. Infrared Phys. Technol. 2016, 76, 353–359. [Google Scholar] [CrossRef]
  141. Li, Y.; Liu, X. Photoluminescence properties of NaY(MoO4)2:Eu3+ phosphor synthesized by microwave assisted hydrothermal method. Mater. Sci. Eng. B Solid-State Mater. Adv. Technol. 2014, 188, 20–25. [Google Scholar] [CrossRef]
  142. Lim, C.S. Highly modulated structure and upconversion photoluminescence properties of PbGd2(MoO4)4:Er3+/Yb3+ phosphors. Mater. Res. Bull. 2016, 75, 211–216. [Google Scholar] [CrossRef]
  143. Chou, W.C.; Chang, C.Y.; Huang, S.C.; Hsieh, T.L.; Chung, S.L. Microwave synthesis of red-emitting Ca2Si5N8:Eu2+ phosphor and its photoluminescence properties for white-LEDs. J. Alloys Compd. 2019, 782, 747–753. [Google Scholar] [CrossRef]
  144. Chen, W.; Liu, M.; Liu, Y.; Lin, Y.; Yu, L.; Hong, J. Synthesis of luminescent SnO2:Eu3+ nanorods via a salt-assisted solution combustion process. J. Nanosci. Nanotechnol. 2014, 14, 5529–5534. [Google Scholar] [CrossRef] [PubMed]
  145. Dominguez, R.D.; Alarcón-Flores, G.; Aguilar-Frutis, M.; Sánchez-Alarcón, R.I.; Falcony, C.; Dorantes-Rosales, H.J.; González-Velázquez, J.L.; Rivas-López, D.I. Effect on the stabilization of the anatase phase and luminescent properties of samarium-doped TiO2 nanocrystals prepared by microwave irradiation. J. Alloys Compd. 2016, 687, 121–129. [Google Scholar] [CrossRef]
  146. Kaszewski, J.; Olszewski, J.; Rosowska, J.; Witkowski, B.; Wachnicki, Ł.; Wenelska, K.; Mijowska, E.; Gajewski, Z.; Godlewski, M.; Godlewski, M.M. HfO2:Eu nanoparticles excited by X-rays and UV-visible radiation used in biological imaging. J. Rare Earths 2019, 37, 1176–1182. [Google Scholar] [CrossRef]
  147. Rahim, S.; Ayob, M.T.M.; Hasim, M.H.; Abdul Rahman, I.; Radiman, S. Physical and optical studies of Gd2O2S:Eu3+ nanophosphors by microwave irradiation and γ-irradiation methods. Luminescence 2019, 34, 699–706. [Google Scholar] [CrossRef]
  148. Bühler, G.; Feldmann, C. Microwave-Assisted Synthesis of Luminescent LaPO4:Ce,Tb Nanocrystals in Ionic Liquids. Angew. Chem. Int. Ed. 2006, 45, 4864–4867. [Google Scholar] [CrossRef]
  149. Li, W.; Lee, J. Microwave-assisted sol-gel synthesis and photoluminescence characterization of LaPO4:Eu3+,Li+ nanophosphors. J. Phys. Chem. C 2008, 112, 11679–11684. [Google Scholar] [CrossRef]
  150. Wang, X.; Wang, Z.; Huang, S.; Shi, X.; Li, J.G. Luminescence responses of CePO4: Tb monospheres toward vitamin C and permanganate. Mater. Res. Bull. 2019, 117, 28–34. [Google Scholar] [CrossRef]
  151. Mokoena, P.P.; Gohain, M.; Bezuidenhoudt, B.C.B.; Swart, H.C.; Ntwaeaborwa, O.M. Luminescent properties and particle morphology of Ca3(PO4)2:Gd3+,Pr3+ phosphor powder prepared by microwave assisted synthesis. J. Lumin. 2014, 155, 288–292. [Google Scholar] [CrossRef]
  152. Rodriguez-Liviano, S.; Aparicio, F.J.; Rojas, T.C.; Hungría, A.B.; Chinchilla, L.E.; Ocaña, M. Microwave-Assisted synthesis and luminescence of mesoporous RE-Doped YPO4 (RE = Eu, Ce, Tb, and Ce + Tb) nanophosphors with lenticular shape. Cryst. Growth Des. 2012, 12, 635–645. [Google Scholar] [CrossRef] [Green Version]
  153. Wang, L.; Xu, M.; Sheng, R.; Liu, L.; Jia, D. Microwave assisted co-precipitation synthesis and photoluminescence characterization of spherical Sr2P2O7:Ce3+,Tb3+ phosphors. J. Alloys Compd. 2013, 579, 343–347. [Google Scholar] [CrossRef]
  154. Zhang, Y.; Xu, J.; Cui, Q.; Yang, B. Eu3+-doped Bi4Si3O12 red phosphor for solid state lighting: Microwave synthesis, characterization, photoluminescence properties and thermal quenching mechanisms. Sci. Rep. 2017, 7, 1–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Birkel, A.; Darago, L.E.; Morrison, A.; Lory, L.; George, N.C.; Mikhailovsky, A.A.; Birkel, C.S.; Seshadri, R. Microwave assisted preparation of Eu2+-doped Åkermanite Ca2MgSi2O7. Solid State Sci. 2012, 14, 739–745. [Google Scholar] [CrossRef]
  156. Yang, H.; Shi, J.; Gong, M.; Cheah, K.W. Synthesis and photoluminescence of Eu3+- or Tb 3+-doped Mg2SiO4 nanoparticles prepared by a combined novel approach. J. Lumin. 2006, 118, 257–264. [Google Scholar] [CrossRef]
  157. Yang, H.; Gong, M.; Liang, H. Fine particle Sr2SiO4:Tb phosphor for the plasma display panel prepared by a combined approach. Mater. Res. Bull. 2010, 45, 805–808. [Google Scholar] [CrossRef]
  158. Sun, J.; Mi, X.; Ma, J.; Su, J.; Xia, W.; Zhang, L.; Bai, Z.; Zhang, X. Rapid synthesis and luminescence properties of Sr3SiO5:Eu2+ phosphors. Funct. Mater. Lett. 2016, 9, 1–5. [Google Scholar] [CrossRef]
  159. Cadiş, A.I.; Muresan, L.E.; Perhaiţa, I.; Popovici, E.J.; Barbu-Tudoran, L.; Indrea, E. Studies regarding ZnS:Mn nanopowders prepared from single source molecular precursor using microwave-assisted decomposition. Mater. Res. Bull. 2016. [Google Scholar] [CrossRef]
  160. Arora, S.; Manoharan, S.S. Structural and photo luminescent properties of uncapped nanocrystalline Cd1-xZnxS solid solutions. Opt. Mater. 2008, 31, 176–180. [Google Scholar] [CrossRef]
  161. Wang, H.; Lu, J.; Wang, R.; Dong, Y.; Ding, L. Synthesis and characterization of the CaTiO3:Eu3+ red phosphor by an optimized microwave-assisted sintering process. Materials 2020, 13, 874. [Google Scholar] [CrossRef] [Green Version]
  162. Liao, J.; Wang, Q.; Lan, L.; Guo, J.; Nie, L.; Liu, S.; Wen, H.R. Microwave hydrothermal synthesis and temperature sensing behavior of Lu2Ti2O7:Yb3+/Er3+ nanophosphors. Curr. Appl. Phys. 2017, 17, 427–432. [Google Scholar] [CrossRef]
  163. Yun, H.-S.; Yun, B.-G.; Lee, H.-M.; Jeong, D.-Y.; Lee, W.-I.; Cho, N.-H. Low-temperature synthesis of nanoscale BaTiO3 powders via microwave-assisted solid-state reaction. SN Appl. Sci. 2019, 1, 1366. [Google Scholar] [CrossRef] [Green Version]
  164. Lim, C.S.; Aleksandrovsky, A.; Molokeev, M.; Oreshonkov, A.; Atuchin, V. Microwave sol-gel synthesis and upconversion photoluminescence properties of CaGd2(WO4)4:Er3+/Yb3+ phosphors with incommensurately modulated structure. J. Solid State Chem. 2015, 228, 160–166. [Google Scholar] [CrossRef]
  165. Wei, B.; Liu, Z.; Xie, C.; Yang, S.; Tang, W.; Gu, A.; Wong, W.T.; Wong, K.L. Fast synthesis of red Li3BaSrLn3(WO4)8:Eu3+ phosphors for white LEDs under near-UV excitation by a microwave-assisted solid state reaction method and photoluminescence studies. J. Mater. Chem. C 2015, 3, 12322–12327. [Google Scholar] [CrossRef]
  166. Fan, M.; Liu, S.; Yang, K.; Guo, J.; Wang, J.; Wang, X.; Liu, Q.; Wei, B. Fast synthesis and energy transfer of the tunable single-phase white-emitting phosphor Li2Gd4(WO4)7:Dy3+, Tm3+ for WLEDs. Ceram. Int. 2020, 46, 6926–6933. [Google Scholar] [CrossRef]
  167. Xu, H.; Xu, K.; Lu, A.; Wang, X.; Hu, J. Microwave hydrothermal synthesis and white up-conversion emission of NaGd(WO4)2:(Yb3+/Tm3+/Ho3+) phosphors. J. Mater. Sci. Mater. Electron. 2015, 26, 3921–3925. [Google Scholar] [CrossRef]
  168. Yang, Y.; Feng, H.; Zhang, X. Microwave heating synthesis and luminescence of NaY(WO4)2:(Ho3+, Yb3+) phosphors. J. Mater. Sci. Mater. Electron. 2015, 26, 229–233. [Google Scholar] [CrossRef]
  169. Liu, Q.; Guo, J.; Fan, M.; Zhang, Q.; Liu, S.; Wong, K.L.; Liu, Z.; Wei, B. Fast synthesis of Dy3+ and Tm3+ co-doped double perovskite NaLaMgWO6: A thermally stable single-phase white-emitting phosphor for WLEDs. J. Mater. Chem. C 2020, 8, 2117–2122. [Google Scholar] [CrossRef]
  170. Mishra, S.; Priyadarshinee, M.; Debnath, A.K.; Muthe, K.P.; Mallick, B.C.; Das, N.; Parhi, P. Rapid microwave assisted hydrothermal synthesis cerium vanadate nanoparticle and its photocatalytic and antibacterial studies. J. Phys. Chem. Solids 2020, 137, 109211. [Google Scholar] [CrossRef]
  171. Jin, Y.; Li, C.; Xu, Z.; Cheng, Z.; Wang, W.; Li, G.; Lin, J. Microwave-assisted hydrothermal synthesis and multicolor tuning luminescence of YPxV1-xO4:Ln3+ (Ln = Eu, Dy, Sm) nanoparticles. Mater. Chem. Phys. 2011, 129, 418–423. [Google Scholar] [CrossRef]
  172. Liu, Y.; Xiong, H.; Zhang, N.; Leng, Z.; Li, R.; Gan, S. Microwave synthesis and luminescent properties of YVO4:Ln3+ (Ln = Eu, Dy and Sm) phosphors with different morphologies. J. Alloys Compd. 2015, 653, 126–134. [Google Scholar] [CrossRef]
Figure 1. Emission spectra of Eu3+ in Eu(TTA)3·(H2O)2 complex (left), and Eu2+-doped BaAl2O4 host (right). Eu3+ sharp lines spectra is attributed to f-f transitions whereas the Eu2+ broad emission band is attributed to f-d transitions. Adapted from references [3,4]. Copyright © 1997 Published by Elsevier B.V.
Figure 1. Emission spectra of Eu3+ in Eu(TTA)3·(H2O)2 complex (left), and Eu2+-doped BaAl2O4 host (right). Eu3+ sharp lines spectra is attributed to f-f transitions whereas the Eu2+ broad emission band is attributed to f-d transitions. Adapted from references [3,4]. Copyright © 1997 Published by Elsevier B.V.
Molecules 26 02882 g001
Figure 2. (A) Downshift mechanism consisting of excitation (1), non-radiative decay (2), and emission (3). (B) Quantum cutting mechanism where the emission of two photons (3,4) occurs after the excitation. (C) Upconversion emission after energy transfer (2) of two lower energy quanta to a higher excited state (3).
Figure 2. (A) Downshift mechanism consisting of excitation (1), non-radiative decay (2), and emission (3). (B) Quantum cutting mechanism where the emission of two photons (3,4) occurs after the excitation. (C) Upconversion emission after energy transfer (2) of two lower energy quanta to a higher excited state (3).
Molecules 26 02882 g002
Figure 3. Representation of the temperature gradient in conventional (left) and microwave-assisted (right) heating of a solid sample.
Figure 3. Representation of the temperature gradient in conventional (left) and microwave-assisted (right) heating of a solid sample.
Molecules 26 02882 g003
Figure 4. Ionic diffusion in a ceramic heated by polarized microwave irradiation (a) and conventional heating (b). Adapted from reference [70]. Reaction rates of NiFe2O4 synthesis under microwave heating (orange) and resistive heating (blue) at 850 °C (c). Adapted from ref. [71]. Copyright © 2005, American Chemical Society.
Figure 4. Ionic diffusion in a ceramic heated by polarized microwave irradiation (a) and conventional heating (b). Adapted from reference [70]. Reaction rates of NiFe2O4 synthesis under microwave heating (orange) and resistive heating (blue) at 850 °C (c). Adapted from ref. [71]. Copyright © 2005, American Chemical Society.
Molecules 26 02882 g004
Figure 5. Fundamental microwave-assisted heating properties that are beneficial in preparing luminescent materials.
Figure 5. Fundamental microwave-assisted heating properties that are beneficial in preparing luminescent materials.
Molecules 26 02882 g005
Figure 6. (a) Microwave-assisted solid-state synthesis setup with hybrid heating using SiC rods as susceptors. (b) MASS synthesis setup using a DMO. Susceptor surrounding the sample absorbs the microwave and initiate the heating [49]. Copyright © 2016, with permission from Elsevier.
Figure 6. (a) Microwave-assisted solid-state synthesis setup with hybrid heating using SiC rods as susceptors. (b) MASS synthesis setup using a DMO. Susceptor surrounding the sample absorbs the microwave and initiate the heating [49]. Copyright © 2016, with permission from Elsevier.
Molecules 26 02882 g006
Figure 7. LIII edge XANES spectra of Pr (left), Eu (middle), and Tb (right) doped Lu2O3. Spectra for reference materials are also provided [62]. Copyright © 2016, American Chemical Society.
Figure 7. LIII edge XANES spectra of Pr (left), Eu (middle), and Tb (right) doped Lu2O3. Spectra for reference materials are also provided [62]. Copyright © 2016, American Chemical Society.
Molecules 26 02882 g007
Figure 8. Microwave-assisted combustion synthesis setup using DMO equipment (A). Synthesis progression at 32 (B), 34 (C), and 44 s (D). Such extensive modification in DMOs needs thorough consideration of the involved hazards, mainly microwave leakage. Reproduced from ref. [85,90] with permission from The Royal Society of Chemistry.
Figure 8. Microwave-assisted combustion synthesis setup using DMO equipment (A). Synthesis progression at 32 (B), 34 (C), and 44 s (D). Such extensive modification in DMOs needs thorough consideration of the involved hazards, mainly microwave leakage. Reproduced from ref. [85,90] with permission from The Royal Society of Chemistry.
Molecules 26 02882 g008
Figure 9. α-Fe2O3 nanorings (a), CuO nanourchins (b), and lamellar Cd(OH)2 (c) obtained by microwave-assisted hydrothermal method. Adapted from ref. [91]. Copyright © 2016, with permission from Elsevier.
Figure 9. α-Fe2O3 nanorings (a), CuO nanourchins (b), and lamellar Cd(OH)2 (c) obtained by microwave-assisted hydrothermal method. Adapted from ref. [91]. Copyright © 2016, with permission from Elsevier.
Molecules 26 02882 g009
Figure 10. Persistent luminescence decay curves of Lu2O3:Pr3+, Hf4+ materials obtained by conventional (blue curve), and microwave-assisted (black curve) solid-state methods, and the PeL decay of the MASS obtained materials without H3BO3 flux (red curve) [62]. Copyright © 2016, American Chemical Society.
Figure 10. Persistent luminescence decay curves of Lu2O3:Pr3+, Hf4+ materials obtained by conventional (blue curve), and microwave-assisted (black curve) solid-state methods, and the PeL decay of the MASS obtained materials without H3BO3 flux (red curve) [62]. Copyright © 2016, American Chemical Society.
Molecules 26 02882 g010
Figure 11. Excitation and emission spectra of Y3Al5O12:Ce3+ (a) and Lu3Al5O12:Ce3+ (b) obtained by both conventional (line) and MASS (dashed line) methods [83]. Copyright © 2012, American Chemical Society.
Figure 11. Excitation and emission spectra of Y3Al5O12:Ce3+ (a) and Lu3Al5O12:Ce3+ (b) obtained by both conventional (line) and MASS (dashed line) methods [83]. Copyright © 2012, American Chemical Society.
Molecules 26 02882 g011
Figure 12. SEM-CL hyperspectral images of the Sr2MgSi2O7:Eu2+(1%), Dy3+(1%) materials obtained by CSS (a,b) and MASS (d,e) method. The numbers in the images (b,e) correspond to the area of the local spectra extracted for the CSS (c) and MASS (f) materials. Reproduced from Ref. [57] with permission from The Royal Society of Chemistry.
Figure 12. SEM-CL hyperspectral images of the Sr2MgSi2O7:Eu2+(1%), Dy3+(1%) materials obtained by CSS (a,b) and MASS (d,e) method. The numbers in the images (b,e) correspond to the area of the local spectra extracted for the CSS (c) and MASS (f) materials. Reproduced from Ref. [57] with permission from The Royal Society of Chemistry.
Molecules 26 02882 g012
Figure 13. Contour plot of the conversion from the hexagonal phase to the monoclinic phase of Eu-doped (Ba1-xSrx)Al2Si2O8 materials. The ramp time and soak time (in minutes) of the microwave reaction are presented on the x and y axes, respectively, while the refined phase fractions are presented on the z axis [112]. Copyright © 2014, with permission from John Wiley and Sons.
Figure 13. Contour plot of the conversion from the hexagonal phase to the monoclinic phase of Eu-doped (Ba1-xSrx)Al2Si2O8 materials. The ramp time and soak time (in minutes) of the microwave reaction are presented on the x and y axes, respectively, while the refined phase fractions are presented on the z axis [112]. Copyright © 2014, with permission from John Wiley and Sons.
Molecules 26 02882 g013
Figure 14. SR-XANES of the sulfur K-edge (top left) and FTIR spectra (top right) of the Gd2O2S materials obtained by the MASS method for different durations of microwave exposure. Schematic representation of oxysulfide formation with the participation of the gases assisting the redox process of sulfur (bottom). Reproduced from Ref. [76] with permission from The Royal Society of Chemistry.
Figure 14. SR-XANES of the sulfur K-edge (top left) and FTIR spectra (top right) of the Gd2O2S materials obtained by the MASS method for different durations of microwave exposure. Schematic representation of oxysulfide formation with the participation of the gases assisting the redox process of sulfur (bottom). Reproduced from Ref. [76] with permission from The Royal Society of Chemistry.
Molecules 26 02882 g014
Figure 15. (a) XRD patterns, (b) excitation spectra, and (c) emission spectra of quinary molybdate phosphors NaLa0.95Pr0.025Dy0.025(MoO4)2 synthesized using conventional and microwave heating. Reproduced from ref. [121] with permission from The Royal Society of Chemistry.
Figure 15. (a) XRD patterns, (b) excitation spectra, and (c) emission spectra of quinary molybdate phosphors NaLa0.95Pr0.025Dy0.025(MoO4)2 synthesized using conventional and microwave heating. Reproduced from ref. [121] with permission from The Royal Society of Chemistry.
Molecules 26 02882 g015
Table 1. Microwave dielectric properties (tan δe and dp) and average heating rate of selected susceptor materials. The values are from the references [76,79,80,81].
Table 1. Microwave dielectric properties (tan δe and dp) and average heating rate of selected susceptor materials. The values are from the references [76,79,80,81].
Materialtan δedp (m)Average Heating Rate (°C/s)
Amorphous Carbon--21
Graphite powder0.36–0.670.013–0.00210
Activated Carbon0.31–0.90.007–0.034.8
Fe3O4 (Magnetite)0.02-7.6
CuO0.0870.212.7
SiC0.3880.0190.3–2.8
Co2O30.0002-7.2
NiO0.001-3.5
Table 2. Dielectric tangent loss (tan δe) of various solvents. Values were obtained at 20 °C and 2.45 GHz microwave frequency [56].
Table 2. Dielectric tangent loss (tan δe) of various solvents. Values were obtained at 20 °C and 2.45 GHz microwave frequency [56].
Solventstan δe
Water0.15
Ethylene glycol1.35
Ethanol0.941
Methanol0.659
DMSO0.825
1-Butanol0.571
DMF0.161
Acetic Acid0.174
Dichloromethane0.042
Table 3. Comparison of microwave-assisted methodologies used to obtain luminescent inorganic materials.
Table 3. Comparison of microwave-assisted methodologies used to obtain luminescent inorganic materials.
Method *AdvantagesDisadvantagesBenefits for Luminescent MaterialsExamples
MASS
  • Energy saving
  • Fast processing times (minutes)
  • Increased reaction rates
  • Increased ionic diffusion
  • High homogeneity
  • Few commercial setups available
  • DMOs can struggle with the reproducibility
  • In modified DMOs safety issues must be addressed
  • High QY
  • High storage capacity
  • Increased lifetimes
  • Less prone to concentration quenching
  • Y3Al5O12:Ce3+ [83]
  • Sr2MgSi2O7:Eu2+, Dy3+ [57]
MAC
  • Fast processing times (seconds)
  • Porous materials
  • Reduced particle sizes
  • Morphology control is difficult
  • Carbon residues
  • In modified DMOs safety issues must be carefully addressed
  • Combination of high homogeneity and moderately small particle sizes
  • Y3Al5O12:Er3+, Yb3+ [85]
  • Gd2O3:Er3+, Yb3+, Zn2+ [90]
MASG
  • Fast processing times (minutes)
  • Nanoparticulate materials
  • Morphology control
  • Carbon residues
  • Expensive and reactive precursors
  • Nanoparticulate materials
  • Possible to use as a coating method
  • ZnGa2O4:Cr3+ [94]
  • M2SiO4:Eu2+ (M:Ca,Ba) [98]
MAH
  • Fast processing times (Seconds to minutes)
  • Nanoparticulate materials
  • Finely controlled morphology
  • DMOs cannot be used due to safety issues
  • Expensive commercial setups
  • Nanoparticles for increased applicability in bioimaging and photodynamic therapy
  • YVO4:Eu3+ [99]
  • ZnWO4:Dy3+ [100]
* Microwave-assisted solid-state (MASS), combustion (MAC), sol-gel (MASG), and hydrothermal (MAH).
Table 4. Inorganic hosts for luminescent materials obtained by microwave-assisted methodologies.
Table 4. Inorganic hosts for luminescent materials obtained by microwave-assisted methodologies.
ClassCompositionMethodSusceptorMW ProgramRef.
AluminateBaMgAl10O17:Eu2+MACUrea and Water15 min @ 850 W[123]
Ca3Al2O6:Eu3+MACSolutionUntil Ignition @ 560 W[124]
Lu3Al5O12:Ce3+MASSCarbon18–25 min@850–1300 W[83]
MgAl2O4MASSCarbon100 min@700 W[75]
SrAl2O4:Eu2+, Dy3+MASSCarbon9 min@960 W + 5 min @480 W[105]
Sr3Al2O6:Eu2+, Dy3+MASSCarbon20 min @ 650 °C[125]
Y3Al5O12:Ce3+MASSCarbon18–25 min@850–1300 W[83]
Y3Al5O12:Ce3+MACSolution5,10,15 min@900 W[126]
Y3Al5O12:Er3+, Yb3+MACSolutionNot provided[85]
ZnAl2O4:Cu+MACSolution1 min@900 W[127]
AluminosilicateAl6Si2O13:Eu3+MASG Solution30–45min@680–800 W[128]
(Ba1-xSrx)1-yEuyAl2Si2O8MASSCarbon10 min@1000 W +
30–90 min@375 W
[112]
(Na,M)8Al6Si6O24(Cl,S)2MASSCarbon18 min@450 W[111]
FluorideKMF3 (M = Zn, Mg, Mn, Co)MASSNone10 min@1100 W[129]
NaYF4:Yb3+,R3+ (R3+: Er, Tm, Yb)MAHSolution5–10 min@60–180 °C[130]
KYF4:R3+ (R: Ce, Tb) MAHSolution4 h@200 °C[131]
LiYF4:R3+ (R: Eu, Tb, Dy)MAHSolution15 min@140 °C[44]
GermanateLa2MgGeO6:Mn4+MASSCarbon40 min@1000 W[122]
Zn2GeO4:Mn2+,Yb3+MAHSolutionNot provided[132]
MolybdateCaMoO4:Eu3+,Dy3+MASSCarbon30 min@560 W[120]
CaGd2(MoO4)4:Ho3+,Yb3+MASGSolution30 min@1250 W[133]
CaMoO4:Er3+, Yb3+MAC Ethylene Glycol23 min@1250 W[134]
CdMoO4:Eu3+MAH Solution10 min @ 180 °C[135]
Cs5Bi(MoO4)4MASS 10min@850 W[136]
Na2MoO4MASSCarbon3 min@840 W (9 cycles)[121]
NaY(MoO4)2:Yb3+, Tm3+MAHSolution1 h@180 °C[137]
NaCaGd(MoO4)3:Ho3+, Yb3+MASGSolution30 min@1250 W[138]
NaCaLa(MoO4)3MASGSolution30min@1250 W [139]
NaCaGd(MoO4)3:Er3+, Yb3+MASGSolution30 min@1250 W[140]
NaY(MoO4)2:Eu3+MAHSolution10 min@170 °C[141]
NaEu(MoO4)2MASSCarbon3 min@960 W (6 cycles)[121]
NaPbLa(MoO4)3:Er3+, Yb3+MASGSolution30 min @ 1250 W[117]
PbGd2(MoO4)4:Er3+, Yb3+MASGSolution30 min@1250 W[142]
NaLa0.95Eu0.05(MoO4)2MASSCarbon3 min@960 W (9 cycles)[121]
NaLa0.95Pr0.025Dy0.025(MoO4)2MASSCarbon3 min@960 W (6 cycles)[121]
NaSrLa(MoO4)3:Er3+, Yb3+MASGSolution30 min @ 1250 W[116]
SrMoO4MAPEthylene
Glycol
20 min@180 W[97]
NitrideCa2Si5N8:Eu2+MASSNone10 min@100–600 W[143]
OxideLu2O3:R3+ (R:Pr, Eu, Tb)MASSCarbon12 min@1000 W + 10 min@900 W[62]
Y2O3:Eu3+US—MAWater60 min@150 W[96]
SnO2:Eu3+MACEthylene GlycolNot Provided[144]
TiO2:Sm3+MAHSolution31 min @ 850 W [145]
HfO2:Eu3+MAHSolution20 min @ 260 °C[146]
OxysulfideR2O2S:Ti,Mg (R:La, Gd, Y)MASSCarbon10 min@900 W + 15 min@800 W[76]
Gd2O2S:Eu3+MASSCarbon2 min@167, 500, and 1000 W[147]
Gd2O2S:Tb3+MASSCarbon10 min@900 W + 15 min@800 W[77]
PhosphateLaPO4:Ce3+, Tb3+MAPIonic Liquid10 s@800 W [148]
LaPO4:Eu3+, Li+MASGGel30 min@80 °C + 30 min@800 °C[149]
CePO4:Tb3+MAPSolution4 min@700 W[150]
Ca3(PO4)2:Gd3+, Pr3+MAHSolution1 h @ 150 °C (300 W)[151]
YPO4:R3+(R:Eu, Ce, Tb)MAHSolutionNot provided[152]
Sr2P2O7:Ce3+, Tb3+MASSSolution15 min @ 1000 °C[153]
SilicateBi4Si3O12:Eu3+MASSNot provided15 min @ 750 °C[154]
Ca2MgSi2O7:Eu2+MASSCarbonNot provided[155]
M2SiO4:Eu2+ (M: Ca, Ba)MASGCarbon10 min + 12 min @ 1000 W[98]
Mg2SiO4:R3+ (R: Eu, Tb)MASGSolution12 min @ 800 W[156]
Na3YSi2O7:Ce3+MASSCarbon8 min@625 W + 15 min@375 W[106]
Sr2SiO4:Tb3+MASGSolution14 min @ 800 W[157]
Sr2MgSi2O7:Eu2+,Dy3+MASSCarbon12 min@1000 W + 10 min@900 W[57]
Sr3SiO5:Eu2+MASSCarbon20 min @3200 W[158]
SulfideZnS:MnMAHSolution10 min@800 W[159]
Zn1-xCdxSMACSolutionNot provided[160]
ZnS:Pb2+MASSNone4 h@ 700 °C[84]
SrS:Eu2+,R3+ (R: Ce, Sm, Er, Dy)MASSCarbon12 min@1000 W + 10 min@900 W[73]
TitanateCaTiO3:Eu3+MASSCarbon10–60 min@528–800 W[161]
Lu2Ti2O7:Yb3+, Er3+MAHSolution1 h@200 °C (100 W)[162]
BaTiO3MASSNone1–5 min@2000 W[163]
TungstateCaGd2(WO4)4:Er3+, Yb3+MASGSolution30 min @ 1250 W[164]
Li3BaSr(La1-xEux)3(WO4)8:Eu3+MASSNot provided20 min@600 °C + 10 min@900 °C[165]
Li2Gd4(WO4)7MASSCarbon10 min@800 °C[166]
NaGd(WO4)2:Yb3+, Tm3+, Ho3+MAHSolution2 h @ 200 °C[167]
ZnWO4:Dy3+MAHSolution1 h@140 °C[100]
Na2WO4MASSCarbon2 min@1080 W (3 cycles)[121]
NaEu(WO4)2MASSCarbon3 min@1080 W (6 cycles)[121]
NaLa0.95Eu0.05(WO4)2MASSCarbon3 min@1080 W (9 cycles)[121]
NaLa0.95Pr0.025Dy0.025(WO4)2MASSCarbon3 min@1080 W (9 cycles)[121]
NaY(WO4)2: Ho3+, Yb3+MAHSolution2 h @ 180 °C[168]
NaLaMgWO6:Dy3+, Tm3+MASSNone10 min @ 900 °C[169]
VanadateCeVO4MAHsolution10 min@600 W[170]
EuVO4MAHSolution15 min@150 °C[99]
Y0.5Eu0.5VO4MAHSolution15 min@150 °C[99]
YPxV1-xO4:R3+ (R: Eu, Dy, Sm) MAHSolution30 min @ 150 °C[171]
YVO4:R3+ (R: Eu, Dy, Sm)MAHSolution20 min @ 270 W[172]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Miranda de Carvalho, J.; Pedroso, C.C.S.; Saula, M.S.d.N.; Felinto, M.C.F.C.; Brito, H.F.d. Microwave-Assisted Preparation of Luminescent Inorganic Materials: A Fast Route to Light Conversion and Storage Phosphors. Molecules 2021, 26, 2882. https://doi.org/10.3390/molecules26102882

AMA Style

Miranda de Carvalho J, Pedroso CCS, Saula MSdN, Felinto MCFC, Brito HFd. Microwave-Assisted Preparation of Luminescent Inorganic Materials: A Fast Route to Light Conversion and Storage Phosphors. Molecules. 2021; 26(10):2882. https://doi.org/10.3390/molecules26102882

Chicago/Turabian Style

Miranda de Carvalho, José, Cássio Cardoso Santos Pedroso, Matheus Salgado de Nichile Saula, Maria Claudia França Cunha Felinto, and Hermi Felinto de Brito. 2021. "Microwave-Assisted Preparation of Luminescent Inorganic Materials: A Fast Route to Light Conversion and Storage Phosphors" Molecules 26, no. 10: 2882. https://doi.org/10.3390/molecules26102882

Article Metrics

Back to TopTop