Next Article in Journal
Discovering New G-Quadruplex DNA Catalysts in Enantioselective Sulfoxidation Reaction
Next Article in Special Issue
Shape-Memory Polymers Hallmarks and Their Biomedical Applications in the Form of Nanofibers
Previous Article in Journal
Detecting Bacterial–Human Lateral Gene Transfer in Chronic Lymphocytic Leukemia
Previous Article in Special Issue
Supported Lipid Bilayer Platform for Characterizing the Membrane-Disruptive Behaviors of Triton X-100 and Potential Detergent Replacements
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Solid-State Preparation of Metal and Metal Oxides Nanostructures and Their Application in Environmental Remediation

by
Carlos Diaz
1,*,
Maria Luisa Valenzuela
2 and
Miguel Á. Laguna-Bercero
3
1
Departamento de Química, Facultad de Ciencias, Universidad de Chile, Las Palmeras 3425, Ñuñoa, Casilla 653, Santiago 7800003, Chile
2
Instituto de Ciencias Químicas Aplicadas, Grupo de Investigación en Energía y Procesos Sustentables, Facultad de Ingeniería, Universidad Autónoma de Chile, Av. El Llano Subercaseaux 2801, Santiago 8900000, Chile
3
Instituto de Nanociencia y Materiales de Aragón (INMA), CSIC-Universidad de Zaragoza C/Pedro Cerbuna 12, 50009 Zaragoza, Spain
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(3), 1093; https://doi.org/10.3390/ijms23031093
Submission received: 6 November 2021 / Revised: 9 December 2021 / Accepted: 13 December 2021 / Published: 20 January 2022
(This article belongs to the Collection Feature Papers in Materials Science)

Abstract

:
Nanomaterials have attracted much attention over the last decades due to their very different properties compared to those of bulk equivalents, such as a large surface-to-volume ratio, the size-dependent optical, physical, and magnetic properties. A number of solution fabrication methods have been developed for the synthesis of metal and metal oxides nanoparticles, but few solid-state methods have been reported. The application of nanostructured materials to electronic solid-state devices or to high-temperature technology requires, however, adequate solid-state methods for obtaining nanostructured materials. In this review, we discuss some of the main current methods of obtaining nanomaterials in solid state, and also we summarize the obtaining of nanomaterials using a new general method in solid state. This new solid-state method to prepare metals and metallic oxides nanostructures start with the preparation of the macromolecular complexes chitosan·Xn and PS-co-4-PVP·MXn as precursors (X = anion accompanying the cationic metal, n = is the subscript, which indicates the number of anions in the formula of the metal salt and PS-co-4-PVP = poly(styrene-co-4-vinylpyridine)). Then, the solid-state pyrolysis under air and at 800 °C affords nanoparticles of M°, MxOy depending on the nature of the metal. Metallic nanoparticles are obtained for noble metals such as Au, while the respective metal oxide is obtained for transition, representative, and lanthanide metals. Size and morphology depend on the nature of the polymer as well as on the spacing of the metals within the polymeric chain. Noticeably in the case of TiO2, anatase or rutile phases can be tuned by the nature of the Ti salts coordinated in the macromolecular polymer. A mechanism for the formation of nanoparticles is outlined on the basis of TG/DSC data. Some applications such as photocatalytic degradation of methylene by different metal oxides obtained by the presented solid-state method are also described. A brief review of the main solid-state methods to prepare nanoparticles is also outlined in the introduction. Some challenges to further development of these materials and methods are finally discussed.

1. Introduction

Over the past few decades, nanoscale particles have caused much interest due to their distinct chemical, physical and biological properties. A variety of nanoparticles (NPs) with various shapes such as spheres, nanotubes, nanohorns, and nanocages, made different materials, from organic dendrimers, liposomes, gold, carbon, semiconductors, silicon to iron oxide, have already been fabricated and explored in many scientific fields, including chemistry, material sciences, physics, medicine, and electronics [1,2,3,4,5]. In this sense, a number of solution methods have been developed for the synthesis of metal and metal oxides nanoparticles [2,3], but few solid states have been reported [6]. The application of nanostructured materials to electronic solid-state devices or to high-temperature technology requires, however, adequate solid-state methods for obtaining nanostructured materials [7,8,9,10,11]. For instance, recent studies reported that the evaporation of solvent to obtain Au nanoparticles in solid state (for adequate incorporation to a solid-state device) results in 3D Au superstructures with properties different to those of Au nanostructures [12]. Thus, solvent-less synthesis of nanostructures is highly significant due to its economical, eco-friendly, and industrially viable nature. Then, the development of new solid-state methods to prepare metallic nanostructured materials is a constant challenge. We have previously informed a new solid-state method to synthesize metallic nanostructures nanomaterials from the pyrolysis of metallic and organometallic derivatives of poly and oligophosphazene under air at 800 °C [13,14,15,16,17]. Nanostructured metals (M), metal oxides (MxOy), and salts (MxPyOz, where P = malonates, succinates, etc.) are obtained, depending on the nature of the metal. Another method when the respective metallic or organometallic derivative is not possible to prepare uses mixtures such as MLn/[NP(O2C12H8)]3 [18,19,20]. In this case, pure phase metallic nanoparticles are obtained. However, in several of these systems, the M or MxOy phase is accompanied by a phosphate phase. These methods have been discussed in detail in several publications [13,14,15,16,17,18,19,20,21,22,23]. In this chapter, we will discuss a novel preparation method of metallic and metal oxides nanostructured particles starting from the macromolecular chitosan·MXn and PS-co-4-PVP·MXn precursor and subsequent solid-state pyrolysis at 800 °C under air (see Figure 1), including their application in environmental remediation. In addition, we will present a brief discussion of some recent solid-state methods to prepare metallic and metal oxides nanoparticles.
In this review, we first present studies of different methods for obtaining solid-state nanomaterials collected from the literature. In addition, a summary of the results obtained in the preparation of nanomaterials using our developed solid-state method is also presented. Finally, results regarding their application in photocatalysis will be discussed for the different nanostructured metal oxides.

2. A General Survey of Methods for Preparation of Metal Oxide Nanoparticles

2.1. The Korgel’s Method

This method consists of the thermolysis of Metal-Thiolate complexes. The products are metal-sulfide nanoparticles with monodispersed and shape distributions. Briefly, the precursor is prepared by a biphasic reaction between a metallic salt and dodecanethiol using sodium octanoate as a phase transfer catalyst to solubilize the metal cationic in a water/chloroform mixture (see Figure 2). The aqueous phase is discarded, and the organic solvent is evaporated. The waxy solid is heated in the air up to 140–240 °C depending on the metal used. Then, the material is redispersed in chloroform to eliminate the rest of the dodecanethiol stabilizer. Different metal-sulfide nanoparticles such as Cu2S [24,25], NiS [26], or Bi2S3 [27], among others, were prepared using this method. In another approach, thermolysis of the related stearate M(OOCC17H33)n affords the respective metal oxides [28].

2.2. The Molecular and Macromolecular Complex Decomposition Method

This method consists simply of the solid-state thermal decomposition of a molecular complex under air at temperatures about 800 °C. Several examples of decomposition of [Ni(en)3](NO3)2 [29], Nickel dimethylglyoximate [30], [bis(2-hydroxy-1-naphtaldehyde) manganese(II)] [31], MnOOH (in presence of Mn3O4) are used to produce NiO and Mn3O4, respectively [32]. In addition, ZnO was obtained from thermolysis of bis(acetylacetonate)Zn(II) [33] and Co3O4 from solid-state pyrolysis of the [Co(NH3)6](NO3)2 complex [34].
A most general solid-state method to prepare Mn, Fe, Co, Ni, Cu, and Zn oxides has been reported from the solid-state decomposition of the metal transition malonates MCH2C2O4·xH2O and transition succinates M(CH2)2C2O4·xH2O [35].
Metal polymeric materials have also been used as precursors of nanostructured metal oxides. For instance, pyrolysis of the [Ru(CO)4]n polymer affords ruthenium nanofibers [36]. Pyrolysis of poly(ferrocenylsilanes) affords α-Fe nanoparticles with 20 ± 5 Å in size [37]. In addition, pyrolysis of poly(ferrocenylsilanes) yields interesting ferromagnetic ceramic composites at 500–1000 °C containing Fe particles in a SiC/C matrix [38].
Metal ligand coordination polymers have also been proposed as useful precursors of nanostructured metal oxides. Pyrolysis at 800 °C of a Ga-acetate polymer, {[Ga(µ-OH)(µ-O2CCH3)2]·HOAc·H2O}n gives rise to nanostructured Ga2O3 [39].
Furthermore, pyrolysis of some miscellaneous solid precursors has been reported. The Os clusters Os3(CO)12 (1), Os4-(µ-H)4(CO)12 (2), and Os6(CO)18 (3), whose schematic representation of their skeletal is shown in Figure 3 [40].
Pyrolysis of these clusters affords Os nanoparticles in the 1 to 10 nm size range. A similar approach was reported using ruthenium clusters, where Ru nanoparticles were obtained.
This method was also proposed for the fabrication of different perovskites. For example, LaMnO3±δ was prepared by thermal treatment of the LaMnOx/C precursors [41]. The solid-state gelation synthesis route was extended to a wide range of interesting perovskite oxides, including LaMnO3, LaFeO3, LaNiO3, LaCoO3, La0.5Sr0.5CoO3, and La0.5Sr0.5Co0.5Fe0.5O3.
Other composites, including perovskites such as CsPbBr3−Al2O3, were synthetized by calcination of the mixture CsBr/PbBr2/AIP = 1:1:30 at 800 °C for 10 min under a nitrogen atmosphere. The as-obtained perovskites possessed a high quantum yield up to 70%, narrow emission line width of 25 nm, and outstanding thermal stability [42].
In addition, perovskites of the type of nanostructured BaTiO3 and SrTiO3 have been prepared on a large scale by a solid state by reaction of the strontium or barium oxalate with TiO2 in anatase phase at 820 °C [43].
Supramolecular metal structures have also been used as useful precursors for nanostructured metal oxides. For instance, Na6[Fe2(µ-O)(µ-CO3)-(chnida)2]·13.5H2O (1; chnida = N-[(3-carboxy-2-oxy-naphthyl)methylene]iminodiacetate), after heating at 1100 °C affords nanoparticles of NaFeO2 [44].
Additional examples of heterostructures include Co3O4/ZnO composites prepared by thermal treatment of the Co3O4/Zn(OH)2 precursor. The as-prepared heterostructure exhibits a high photocatalytic activity toward Rhodamine B dye higher than ZnO [45]. In a similar way, Fe3O4@M (where M = Au, Ag, and Au-Ag alloy) core-shell nanostructures were synthesized to gram scale in the laboratory conditions by a similar high-temperature solid-state method [46]. The method consists of the thermal treatment of the solid-state mixture of the respective metallic salts with Fe3O4. The as-obtained Fe3O4@M nanocomposites exhibited catalytic activity in the obtention of H2 from NH3BH3 and NaBH4.
Thermal decomposition of organometallic precursors is another usual solid-state route to prepare nanostructured metal oxides. For instance, thermal treatment of the ferrocene carboxaldehide gives Fe2O3 nanoparticles (hematite phase) with an average size of 5 nm [47].
Another synthesis strategy arises from the solid-state thermolysis of the metal organic framework (MOFs) [48]: Zn(ADA)(4,4′-bipy)0.5, [Mn2(hfbba)2(3-mepy)]●(H2O)] [Mg3(O2CH)6I[NH(CH3)2]0.5], [Cd(ADA)(4,4′-bipy)0.5]●(DMF)], [Cd(tdc)(bpy)(H2O)]n [Cu3(TMA)2(H2O)3]n and [Co6(BTC)2(HCOO)6(DMF)6] affords the nanostructured of Cu/CuO, Co/Co3O4, ZnO, Mn2O3, MgO, CdS/CdO. These nanoparticles dispersed in a carbon matrix showed promising H2 and CO2 adsorption properties depending on the environment used for the thermolysis of MOFs.
In a similar manner, additional interesting precursors were the mixture of glucose-urea-transition metal, which, after pyrolysis, gave rise to a series of 2D porous metal oxides La0.5Sr0.5Co0.8Fe0.2O3, Co3O4, NiCo2O4, and RuO2 and 1D nanowire Ba0.5Sr0.5Co0.8Fe0.2O3, which were obtained by calcination in the air [49]. The precursors were prepared first, synthesizing the glucose-urea deep eutectic solvent on which the metallic precursors were added. The as-prepared materials exhibit high activity for the electrochemical oxygen evolution.
Finally, polynuclear clusters have also been used in the preparation of Cr2O3 nanoparticles using a solid-state method starting from direct thermal decomposition of [Cr3O(CH3CO2)6(H2O)3]NO3●CH3COOH ([Cr3O]) in Ar atmosphere [50]. The nanoplates embedded in carbon show an efficient enhanced electrochemical performance.
Another similar example is the solid-state preparation of CoFe2O4/C from thermal treatment of a heterometallic trinuclear [CoFe2O(CH3COO)6(H2O)3·2H2O] [51] complex, showing an average particle size of 50 nm coated with carbon on the surface.

2.3. A Novel Solid-State Approximation

Recently, we have proposed a novel solid-state approximation including two steps [52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67]:
  • The preparation of macromolecular precursors of general formula chitosan●MXn and PS-co-4-PVP·MXn;
  • Pyrolysis of the chitosan·MXn and PS-co-4-PVP●MXn.
The preparation of the macromolecular chitosan●MXn and PS-co-4-PVP·MXn precursors is performed by a simple coordination reaction using dichloromethane as solvent at room temperature (see Equations (1) and (2)).
M x n + C h i t o s a n C H 2 C l 2 C h i t o s a n · M x n
M x n + P S c o 4 P V P C H 2 C l 2 P S c o 4 P V P · M x n
Due to the large coordination site of the polymer (e.g., about 350 for the chitosan Mw = 61.000) and the insolubility of some reactants, the reaction is slow. Ensuring a high percentage of coordination takes about two weeks. However, the insoluble product is easily separable by decantation and dried in vacuum. Products are stable, solid, and with the characteristic color of the precursor metallic salt (e.g., white for Ti, green for Cr and Ni, etc.).
Pyrolysis of the solid chitosan·MXn and PS-co-4-PVP·MXn precursors under air at 800 °C, usually in a ceramic crucible, affords the nucleation of MXOY metal oxides. The obtained results will be shown in order of the periodic table.

2.3.1. Transition Metal (First Row)

Titanium

Titanium dioxide (TiO2) is a well-known semiconductor, which can crystallize in eight different polymorphic forms [67], e.g., rutile, anatase, brookite, TiO2-B (bronze). In particular, anatase is known to be a potential solar-driven photocatalyst active for the photodegradation of various dye contaminants [68]. For this study, we selected the following precursors: (Chitosan)·(Cp2TiCl2) (I), (PS-co-4-PVP)·(Cp2TiCl2) (II), (chitosan)·(TiOSO4) (III) and (PS-co-4-PVP)·(TiOSO4) (IV) and (chitosan)·(Ti(acac)2) (V) and (PS-co-4-PVP)·(Ti(acac)2) (VI). As shown in Table 1, the obtained polymorph of the pyrolytic products depends on the nature of both the polymer and the metallic precursor. It is noteworthy that the chitosan polymer affords a pure rutile phase using Cp2TiCl as metallic salt joined to the polymeric chain and pure anatase using TiO+2 joined to the polymeric chain. The polymer PS-co-4-PVP induces a mixture of rutile and anatase phases. A summary of the different products obtained using the different precursors at several temperatures, average sizes, band gaps, and morphologies are displayed in Table 1.
Interestingly, the product from the (TiOSO4)•(Chitosan) precursor at all temperatures affords a pure anatase phase. In contrast, the solution synthesis methods almost always give phase mixtures such as anatase/rutile or anatase/brookite. As it will be later discussed in detail, the photocatalytic activity toward blue methylene was tested for TiO2 obtained from all precursors. The best photocatalyst, the anatase obtained from the (chitosan)•(TiOSO4) precursor at 800 °C, achieved a 98% discoloration rate in only 25 min when the pH of the solution was 9.5, improving the efficiency of the standard Degussa P25 photocatalyst without the addition of other phases or dopants.

Vanadium

Owing to its different oxidation states, vanadium has several metal oxides being the most common VO, V2O3, VO2, and V2O5 [69]. The most common morphology for V2O5 is the nucleation in lamellar form. Different morphologies as bundles with spindle-like morphologies [70], V2O5 macro-plates, nanoribbons, nanowires, or nanorods have been reported [71]. For our study, we have selected the chitosan•(VCl3)n and PS-co-4-PVP•(VCl3)n precursors. The same pyrolytic V2O5 product was obtained for precursors in molar ratios 1:1 and 1:5 [72]. The smallest nanoparticles were obtained for the chitosan precursor in a 1:5 ratio (see Figure 4), where nanoparticles as small as 8 nm were observed.
Although several solution methods to prepare nanoparticles of V2O5 [69,70,71] have been described, only a few solid-state methods have been reported [73,74,75]. For example, solid-state thermal decomposition of [NH4V3(OH)6(SO4)2] affords single-crystalline V2O5 nanoparticles, used as cathode material for lithium-ion batteries [73]. In addition, pyrolysis of the solid [N3P3(OC6H5)5(OC5H4N·Cp2VCl][PF6] affords mixtures of V2O5/VO(PO3)2 [74].

Chromium, Molybdenum, and Tungsten

The most common oxides of the VI group are Cr2O3, MoO3 and WO3. These metal oxides can be obtained using the same pyrolitic method, from the respective (chitosan)·(CrCl3)x and PS-co-4-PVP·(CrCl3)x, PS-co-4-PVP●(MoCl4)n and chitosan●(MoCl4)n, and from (chitosan)●WCl4 and PS-co-4-VP·WCl4 precursors [76]. For Cr2O3 nanoparticles, the size can be controlled by the metal/polymer ratio decreasing in the order 1:1 > 1:5 > 1:10.

Manganese

For this study, we selected the precursors chitosan·(MnCl2)n and PS-co-4-PVP·(MnCl2)n. XRD analysis clearly shows the presence of Mn2O3 [77]. Figure 5 shows SEM analysis, indicating the presence of dense grains, some of them fused in 3D arrangement. EDS analysis confirmed the formation of manganese oxide.

Iron

Among the different iron oxides, one of the most common is Fe2O3 hematite [78,79]. For the study with Fe, the following macromolecular complexes were prepared: Chitosan·(FeCl2)n, chitosan·(FeCl3)n, PS-co-4-PVP·(FeCl2)n and PS-co-4-PVP·(FeCl3)n with molar ratios 1:1, 1:5, and 1:10. A complete study was performed using the following parameters: nature of the polymer, oxidation state of the iron salts, and the metal/polymer ratio. In all the cases, Fe2O3 (hematite) was obtained. A representative XRD for the pyrolytic product from chitosan·(FeCl3)n 1:1, their SEM and TEM images are shown in Figure 6. XRD data were indexed according to the hematite structure [78,79,80,81,82,83]. SEM images (Figure 6b–d) exhibit varied irregular shapes, typical of solid-state methods, but show the correct Fe and O content atoms as shown in the EDS analysis (Figure 6e). Consistently, their TEM analysis exhibits agglomerates containing the smallest nanostructures joined in linear dispositions (Figure 6f–h). Chitosan with the FeCl2 salt induces the smallest Fe2O3 nanoparticle size, while for the PS-PVP polymer, the smallest nanoparticles were induced with FeCl3. The 1:1 molar ratio precursor also exhibits the smallest nanoparticles for both polymers.

Cobalt

Among the different cobalt oxides, Co3O4 and CoO [84,85,86,87,88,89] are the most common, being Co3O4, for example, used as a supercapacitor electrode material [86]. In this case, chitosan·(CoCl2)n and PS-co-4-PVP·(CoCl2)n precursors were selected. XRD analysis shows the presence of Co3O4 pure phase [84,85], as shown in Figure 7a. SEM images indicate after pyrolysis a dense morphology composed of joined grains to form a 3D arrangement. A similar thermal synthetic approach has been reported [87,88,89]. As normally found in solid-state methods, morphology with different and varied shapes and sizes are observed. In addition, a solid-state route starting from Co(II) salts has also been reported [84].

Nickel

For this study, we have selected the chitosan·(NiCl2)n and PS-co-4-PVP·(NiCl2)n precursors. XRD analysis clearly shows the presence of NiO pure phase. SEM image exhibits a porous morphology, as shown in Figure 8a. EDS analysis corroborates the presence of Ni and O. For the PS-co-4-PVP·(NiCl2)n precursor, a more porous morphology was observed (Figure 8c). Their EDS analysis also indicates the presence of Ni and O, see Figure 8d.
NiO is a p-type semiconductor with a band-gap in the range 3.6–4 eV [60]. For their preparation, several solution methods [60] have been reported. In addition, a solid-state thermal method starting from [Ni(en)3][NO3]2 [29] has been described. The method reported here allowed us to prepare NiO from inexpensive and commercially available polymer and Ni salts. Using chitosan·(NiCl2·H2O)x and PS-co-4-PVP·(NiCl2·6H2O)x precursors [65] and by thermal treatment at 800 °C under air, pure NiO phase was obtained. The band-gap of the as-prepared NiO is 4.15 eV for the one obtained from both Chitosan and PS-co-4-PVP precursors [60]. For the semiconductor metal oxides, their band-gap value dictates their photocatalytic activity [89,90,91,92]. In fact, degradation of methylene blue was 71% and 68% using NiO obtained from Chitosan and PS-co-4-PVP precursors [60].

2.3.2. Noble and Precious Metals

Noble metals such as Au, Pt, and Ag and metal oxides of the precious metal such as Ir, Rh, and Re can also be obtained using the same method. All these metal and metallic nanoparticles can be obtained by pyrolysis of the respective macromolecular chitosan·(MLx) and PS-co-4-PVP·(MLx) precursors using the appropriate metallic salts MLx = AuCl, AuCl3, AuC6F5, Au(PPh3)Cl for Au, Ag(CF3SO3)2 for Ag, PtCl2 for Pt, IrCl3 for Ir, RhCl3 for Rh and ReCl3 for Re.

Gold

The facile reaction of the Au metallic salts = AuCl, AuCl3, AuC6F5, Au(PPh3)Cl with the PS-4-co-PVP polymer leads to the luminescent macromolecular complexes PS-4-co-PVP·AuCl3, PS-4-co-PVP·AuCl, PS-4-co-PVP·AuC6F5, and PS-4-co-PVP·Au(PPh3)Cl. The observed luminescence for PS-4-co-PVP·AuC6F5 with a maximum around 550 nm, as shown in Figure 9, arises probably from the Au(I)-Au(I) interactions [93]. Interestingly, the macromolecular complex PS-4-co-PVP●(AuCl3)n does not present this luminescent behavior.
The pyrolysis of all these macromolecular precursors affords metallic Au° foams with pore size depending on the nature of the Au salt joined to the polymer chain [94] (Figure 10).
Metallic sponges (or foams) of macroporous metals have attracted great attention due to their unusual and peculiar properties such as mechanical strength and stiffness [95]. These materials can be prepared by several solution methods as dealloying of M/M′ alloys and by forming a metal organic composite and eliminating the organic part by dissolution or calcination. However, no solid-state approximations have been reported. The method described here could be a novel and reliable way to prepare metallic foams from different noble metals.

Silver

Several solution methods for preparing Ag° nanoparticles have been reported but relatively scarce from solid-state methods [96,97,98]. For Ag° containing precursors as Ag(CF3SO3)·Chitosan, Ag was obtained as shown in the XRD pattern [94]. Their morphology indicates a foam-like shape. Finally, EDS analysis confirmed the presence of Ag as a pure single phase.

Platinum

Nanostructured Pt nanoparticles are very important, for example, in the catalysis of fuel cells, sensors, and the petroleum and automotive industries due to their high catalytic activity and stability [98].
Pt nanoparticles were obtained from the chitosan·(PtCl2)n and PS-co-4-PVP·(PtCl2)n precursors [99]. As shown in Figure 11, the obtained Pt nanoparticles exhibit varied shapes and sizes, some of them with the typical truncated octahedron [99,100]. This also can be viewed as an image of a cubohedral structure viewed along the [101] zone axis.
The smallest particle size (6 nm) was obtained from the pyrolytic product from the chitosan·(PtCl2)n precursor in a 1:1 molar ratio. In spite of numerous preparation methods of Pt nanoparticles reported [101,102,103], few solid-state methods have appeared.
Another interesting aspect is the “foam-like” morphology observed for pyrolytic products from the precursors 1:5 ratio PS-co-4-PVP·(PtCl2)n. Few metallic Pt sponges materials have been reported, being this the only solid-state route for this type of material [95].

Iridium

Among the metals of the periodic table, the precious as Ir is one of the most catalytically active [104]. Their activity is hugely enhanced at the nano level [105,106,107]. This metal, as well as its metal oxides, exhibits a high catalytic activity [104,106]. Although isolated solution preparation methods for nanostructured Ir oxides are well documented [108,109,110,111,112,113,114], no solid-state general methods to prepare IrO2 nanostructured have been shown. IrO2 is generally prepared from an Ir salt. The relative fraction of IrO2/Ir depends on the temperature, producing Ir2O3 at temperatures between 250 and 400 °C, and then obtaining pure IrO2 at temperatures above 600 °C [108,109,110,111,112,113,114]. Using our method, we have obtained a unique nanostructured phase of IrO2 [63]. The IrO2 nanoparticles were prepared by thermal treatment of the macromolecular chitosan·(IrCl3)X and PSP-4-PVP·(IrCl3)X precursors. The nature of the polymeric precursor is acting as a solid-state template and influences the size of the iridium dioxide but not significantly the morphology, and the obtained IrO2 nanoparticles are about 15 nm.

Rhodium

Rhodium is another catalytically active precious metal [115]. Their activity is also hugely enhanced at the nano level [115,116]. Among these various noble metals, rhodium plays an important role in various catalytic applications [117,118]. However, the catalytic mechanism of rhodium-containing materials is still elusive. Recent investigations suggest that the active centers could be rhodium oxide rather than rhodium [118,119,120,121,122]. The most common rhodium oxides are Rh2O3 and RhO2. Although these rhodium oxides have a wide range of applications in catalysis, scarce preparation methods of nanostructured Rh2O3 and RhO2 have been reported, and their morphological and size control is vaguely known [7,8,9,118,119,120,121,122].
A nanostructured Rh/RhO2 mixture was easily obtained by thermal treatment of the macromolecular chitosan·(RhCl3)X precursor, while the pure Rh2O3 was obtained from pyrolysis of the PSP-4-PVP·(RhCl3)X precursor [64]. The nature of the polymeric precursor acting as a solid-state template does not significantly influence the morphology of the Rh and their metal oxide. The average size of the as-obtained products is in the range of 20 nm.

2.3.3. Representative Metals

SnO2 and ZnO are two of the most used materials in sensors [123,124,125]. SnO2 is a wide band-gap n-type semiconductor with great importance in several technological applications such as gas sensing, Li-ion batteries, and solar cells [124]. As a sensor, its main application is in H2 and CO detection. In addition, nanostructured ZnO is one of the most promising nanomaterials for sensors due to its biocompatibility, chemical and photochemical stability, high specific surface area, optical transparency, electrochemical activity and high electron mobility. ZnO has been employed for the detection of biological molecules [126].
SnO2 and ZnO were prepared by pyrolysis of macromolecular complexes: PS-co-4-PVP·(SnCl2)n and PS-co-4-PVP·(ZnCl2)n in several molar ratios under air at 800 °C [53]. For ZnO agglomerates, the respective hexagonal and cubic structures were observed with typical sizes of 30–50 nm for a precursor mixture ratio of 1:1.
Figure 12 shows, for instance, the SEM image for the pyrolytic precursor from PS-co-4-PV·(ZnCl2)n in 1:5 ratio where zones with “metallic foams” as well as cubic morphologies were observed.
For the SnO2 semiconductor, the oxide was prepared from the PS-co-4-PVP·(SnCl2)n precursor. The nanostructured SnO2 exhibits morphologies and particle sizes depending on the molar ratio of the SnCl2:PS-co-4-PVP. When a larger weight fraction of the inorganic salt is used in the precursor (1:1), larger crystalline crystals are found for each oxide. As shown in Figure 13, the SEM image indicates an irregular morphology, including some “foam” shapes. EDS analysis confirmed the presence of Sn and O atoms.
Nanostructured ZnO was also obtained from the macromolecular chitosan·(ZnCl2) and PSP-co-4-PVP·(SnCl2)n complexes [29].

2.3.4. Rare Metals

Rare earth compounds have drawn attention due to their unique properties and promising application in, for example, UV-shielding, luminescent display, optical communications, biochemical probes, and medical diagnostic [127,128,129,130]. Among the rare earth compounds, metal oxides showed a variety of attractive features for applications in several fields of technology. However, few general preparation methods have been reported [131]. Using the above-described method, we can obtain rare earth oxide nanostructured compounds, although in some cases, the lanthanum oxyhalides were also obtained instead of the expected oxide Ln2O3. LnOX (Ln = lanthanide element, X = halide) compounds have unique and excellent characteristics in electrical, magnetic, optical and luminescent properties [132,133].
Pyrolysis products of the macromolecular PSP-co-4-PV·(MLn)n and chitosan●(MLn)n (M = lanthanide metal) complexes afford products of composition M2O3 or MOCl depending on the nature of the MLn salt. When MLn is MCl3, the product is the oxychloride MOCl, while when MLn is M(NO3)3 or M2(SO4)2, the respective M2O3 is obtained. For instance, pyrolysis of the macromolecular PSP-co-4-PVP·(Ce(NO3)3)n precursor gives rise to CeO2. The different obtained materials were identified by XRD. TEM shows the typical arrangements of nanoparticles somewhat agglomerated, as is displayed in Figure 14a. HRTEM images in Figure 14b confirm that the method allowed agglomerates of single-crystal nanoparticles of CeO2.
On the other hand, pyrolysis of chitosan●NdCl3 and PSP-co-4-PVP·NdCl3 give rise to NdOCl in both cases, as confirmed by XRD [60]. In some cases, a mixture with Nd2O3 oxide was also observed. SEM image exhibits a 3-D metallic foam morphology for the pyrolytic products from chitosan·NdCl3. EDS confirmed the presence of Nd, O, and Cl in agreement with the proposed formula. TEM image displayed irregular shapes with varied sizes [60], as shown in Figure 15c.
It is remarkable that Eu+3 doped NdOCl/Eu2O3 materials obtained from pyrolysis of chitosan●NdCl3/EuCl3 exhibit a 3D metallic foam structure (Figure 15) [57]. Interestingly, we observed the presence of Eu, in addition to the expected Nd, O, and Cl atoms, in the respective EDS (Figure 15b). The solid-state luminescence of this system shows the main emission line around 566 nm assigned to the 5D02F2 transition [57].

2.3.5. Actinides

Among the different actinide oxides, thoria is an important and promising material used as a ceramic catalyst sensor in solid electrolytes, catalysis, and optical materials, as well as in the traditional nuclear industry [134,135,136,137,138]. In spite of this, few papers related to the preparation and properties of nanostructured ThO2 have appeared. We have prepared nanostructured ThO2 from the chitosan·Th(NO3)4 and PS-co-4-PVP·Th(NO3)4 precursors [66]. The morphology and the average size of the as-obtained ThO2 depend on the Chitosan and PS-co-4-PVP polymer forming the precursor. A total of 50 and 40 nm average sizes were observed from the Chitosan and PS-co-4-PVP polymer precursors. The as-obtained thoria exhibits the expected luminescence with a dependence on their intensity emission maxima and the nature of the precursor polymer.

3. Incorporation of Metallic and Metal Oxides into Solid Matrix

Various practical applications, for instance, catalysis, involving solid-state devices are formed by nanoparticles and/or nanostructured inside a solid matrix as SiO2, TiO2, Al2O3, glasses, and so on [139,140]. We have designed a solid-state methodology to prepare different composites: M/M′xO′y and MxOy/M′xO′y with M′xO′y solid matrix, i.e., SiO2, TiO2, Al2O3, under air thermal treatment of the chitosan·MLn//M′xO′y and PS-co-4-PVP·MLn//M′xO′ precursors. Using this synthetic methodology, we have been able to prepare several metallic nanoparticles as well as metal oxides included in solid matrices.
As known in catalysis, the inclusion of M°, as well as MxOy inside solid matrices, induces stability of the catalytic material as well as a greater surface area of the catalyst within the solid matrix [139,140]. Au° and the bimetallic Au°/Ag° were incorporated in silica to give Au°//SiO2 [94] and Au°/Ag°//SiO2 [56].
Nanoparticles of about 5 nm can be observed in Figure 16. The distribution of the Au° nanoparticles inside silica was examined by SEM-EDS mapping, as seen in Figure 17.
From this figure, regular distribution of the Au° nanoparticles inside silica was observed. In addition, the inclusion of the bimetallic Au°/Ag° nanoparticles was performed by pyrolysis under air at 800 °C of the PSP-4-PVP•(AuCl3/AgSO3CF3)n·SiO2 and chitosan·(AuCl3/AgSO3CF3)n·SiO2 precursors. The relative distribution of the Au/Ag bimetallic nanoparticles inside silica was observed by the EDS mapping shown in Figure 18.
The inclusion of the Ag° nanoparticles inside SiO2 was also made by pyrolysis of the respective (PS-co-4-PVP)·(AgNO3)n•(SiO2)n and chitosan·(AgNO3)n•(SiO2)n precursors [94]. Well dispersed Ag nanoparticles inside SiO2 were observed for the Ag/SiO2 composites obtained from chitosan•(AgNO3)n•(SiO2)m and (PS-co-4-PVP)·(AgNO3)n•(SiO2)m, with particle sizes of 5 and 6 nm, respectively.
Furthermore, ZnO and SnO2 were also included in silica [58]. Pyrolysis of the ZnCl2·chitosan·SiO2 and SnCl2·chitosan·SiO2 precursors at 800 °C under air afford mixtures of Zn2SiO4 and SiO2, and pure SnO2, respectively. SnO2 nanoparticles are regularly distributed inside the silica matrix.
The inclusion of IrO2 into SiO2 was performed using a combined solution of the chitosan and PVP precursors using the sol-gel method [63]. Subsequent pyrolysis of the isolated solid-state chitosan·(IrCl3)x(SiO2)y and PSP-4-PVP·(IrCl3)x(SiO2)y give rise to the IrO2//SiO2 nanocomposites. The IrO2 particles are distributed uniformly inside the matrix of SiO2, leading to stable porous materials appropriate for high-temperature catalytic applications.
The same procedures were used for Re and Th. The inclusion of ReO3 into SiO2 was performed using a combined solution of the chitosan and PSP-4-PVP precursors by the sol-gel method [62]. Subsequent pyrolysis of the solid chitosan·(ReCl3)X(SiO2)y and PSP-4-PVP·(ReCl3)X·(SiO2)y precursors give rise to the ReO3//SiO2 nanocomposites. The as-obtained ReO3 nanoparticles inside SiO2 are as small as 1 nm. ReO3 nanoparticles are distributed uniformly inside the SiO2 matrix, leading to stable semiporous materials suitable for high-temperature catalytic applications. The inclusion of ThO2 inside SiO2 and TiO2 was achieved through a similar solid-state method [140,141,142]. The ThO2/SiO2 composites were prepared by pyrolysis at 800 °C under air of the chitosan●Th(NO3)4//SiO2 and PS-co-4-PVP·Th(NO3)4//SiO2 precursors [66]. On the other hand, ThO2/TiO2 composites were similarly prepared by pyrolysis at 800 °C chitosan·Th(NO3)4//TiO2 and PS-co-4-PVP·Th(NO3)4//TiO2 precursors. ThO2 particles exhibit a suitable dispersion inside the silica showing sizes of 250 nm and 950 nm depending on the chitosan or PS-co-4-PVP polymer precursors, respectively. SEM-EDS mapping analysis shows a regular dispersion of the thoria into the SiO2 and TiO2 matrices. The luminescent properties of the ThO2/SiO2 and ThO2/TiO2 composites show a dependence of their luminescence intensity, being the most intense with the TiO2 matrix.
On the other hand, the inclusion of NiO inside the SiO2, TiO2, Al2O3, Na4.2Ca2.8(Si6O18) matrices was also performed by solid-state under-air pyrolysis of the: Chitosan·(NiCl2·6H2O)x//SiO2, PS-co-4-PVP·(NiCl2)x//SiO2, chitosan·(NiCl2·6H2O)x//TiO2, PS-co-4-PVP(NiCl2)x//TiO2, chitosan(NiCl2·6H2O)x//Al2O3, PS-co-4-PVP·(NiCl2)x/Al2O3, chitosan(NiCl2·6H2O)x//NiO/Na4.2Ca2.8(Si6O18) and PS-co-4-PVP(NiCl2)x//NiO/Na4.2Ca2.8(Si6O18) precursors [65]. The new composites were characterized by XRD, SEM/EDS, TEM, and HR-TEM. The size of the NiO nanoparticles obtained from the PSP-4-PVP precursors inside the different matrices follows the order of SiO2 > TiO2 > Al2O3. However, NiO nanoparticles obtained from the chitosan precursor do not present an effect on the particle size. It was found that the matrices (SiO2, TiO2, Al2O3, and Na4.2Ca2.8(Si6O18)) have a medium effect on the band-gap energy and also on the photocatalytic methylene blue degradation.

4. Photocatalytic Applications

Industrial plants generate increasing amounts of wastewaters, which often causes severe environmental problems [141,142]. Wastewaters produced in many industrial processes typically contain organic compounds that are toxic and not amenable to direct biological treatments [142]. There are huge numbers of different types of organic pollutants, including dyes, phenols, biphenyls pesticides, fertilizers, hydrocarbons, plasticizers, detergents, oils, greases, pharmaceuticals, proteins, carbohydrates, and so on [143]. Therefore, there is a great need to develop an efficient and cost-effective technique to reduce the concentration of organic pollutants before releasing the wastewaters into the aquatic environment. Currently, industrially available wastewaters treatment technologies such as adsorption and coagulation merely concentrate or separate these pollutants from water but do not completely “eliminate” or “destroy” them into biodegradable or less toxic organic compounds [144]. Other water treatments methods, such as chemical and membrane technologies, usually involve high operating costs and sometimes generate other toxic secondary pollutants [145]. Among the various physical, chemical, and biological technologies used in pollution control, including biological technologies, advanced oxidation processes such as photocatalysis are being increasingly adopted in the destruction of the organic contaminant due to their high efficiency, simplicity, suitable reproducibility, and ease of handling [146]. Heterogeneous photocatalysis possesses some critical advantages that have feasible applications in wastewater treatments, including:
(i)
Ambient operating temperatures and pressure;
(ii)
Complete mineralization of contaminants and their intermediates compounds without leaving secondary pollutants;
(iii)
Low operating costs [146].
Among the most used photocatalyst are the nanostructured metal oxides [10,11]. However, their current preparation involves mainly in-solution methods [147,148,149,150,151,152,153,154], which present some problems in the isolation of the solid by elimination of the solvent as well as the elimination of the template and of the stabilizer [151].
Among the main applications of nanostructured metal oxides, the photocatalytic degradation of organic pollutants lies in the field of environmental remediation. The main characteristics that a suitable metal oxide photocatalytic system must include [145]:
  • An adequate band-gap;
  • Suitable morphology;
  • High surface area;
  • Stability and reusability.
Semiconductor metal oxides having a band-gap near 3.2 eV are UV light active, while semiconductor metal oxides with a band-gap near 2.7 eV are visible light active [142]. Metal oxides exhibiting these features, such as vanadium, chromium, titanium, zinc, tin, and cerium, follow similar primary photocatalytic processes such as light absorption, which induce a charge separation process with the consequent formation of positive holes that are able to oxidize organic substrates. In this process, a metal oxide is activated by either UV light, visible light, or a combination of both, and photoexcited electrons are promoted from the valence band to the conduction bands, forming an electron/hole pair (e/h+). The photogenerated pair (e/h+) is able to reduce and/or oxidize a compound adsorbed on the photocatalyst surface. A schematic representation of these processes is shown in Figure 19.
The photocatalytic activity of metal oxides comes from two sources [148]:
  • Generation of •OH radicals by oxidation of OH anions;
  • Generation of O2 radicals by reduction in O2.
Semiconductor nanostructured metal oxides have been widely used in photocatalytic redox processes because of their electronic configuration of the filled valence band (VB) and empty conduction band (CB). When exposed to a photon with energy exceeding the band-gap, hν > Eg, it generates an electron-hole pair with one electron in VB pumped into CB, leaving a hole behind in VB. The generated holes in VB are of great oxidation capability, while the electrons in the CB have high reducing potential.
These highly reactive electrons and holes participate in photocatalytic organic degradation.
As mentioned above, the factors that are important for an efficient photocatalyst include an adequate band-gap, suitable morphology, high surface area, stability, and reusability. The achievement of these characteristics for a given nanostructured metal oxide will depend on its preparation method. For instance, TiO2 is one of the most used and efficient metal oxides for photocatalytic degradation of several organic dye pollutants [145]. However, its relative efficiency depends on the preparation method, which in turn determinates the band-gap, the morphology, the surface area, and their stability and reusability. In this context, our solid-state method could afford nanostructured metal oxides that can easily exhibit the above characteristics, yielding an efficient photocatalyst for the degradation of organic pollutants.
One of the advantages of the metal oxides obtained by the solid state described above is that they can be used directly in photocatalytic heterogeneous catalysis.
For all TiO2 products described in Table 1, the most effective degradant of methylene blue was the anatase obtained from the precursor (chitosan)·(TiOSO4) at 800 °C [59]. This material achieved a 98% discoloration rate in only 25 min when the pH of the solution was 9.5, improving the efficiency of the standard photocatalyst Degussa P25 without the addition of other phases or dopants. Figure 20a shows the c/c0 vs. irradiation time for the TiO2 from (chitosan)·(TiOSO4) at several temperatures. At 800 °C, 86.5% discoloration rate in 25 min was observed. Optimization of the pH shows a 98% discoloration at pH 9.5, as shown in Figure 20b. From all different known TiO2 materials, the obtained using our solid-state method is one of the most efficient toward methylene blue degradation [59].
Another photocatalytic-assayed system was hematite-Fe2O3 [54]. The nanoparticulate material obtained from chitosan·(FeCl2)y 1:1 under the simulated sunlight (full visible spectrum) irradiation provides high rate degradation of MB by 73% in 60 min and >94% after 150 min, measured at 655 nm, as seen in Figure 21. The high photocatalytic efficiency can be due, in part, to the porous morphology of the hematite-Fe2O3 [54].
On the other hand, NiO as well as their NiO/SiO2, NiO/TiO2, NiO/Al2O3 nanocomposites presenting band-gap values in the range 5.0–5.6 eV, see Table 2, predicts that they can be used as appropriate photocatalyst using UV irradiation [65]. In fact, NiO and their NiO/SiO2, NiO/TiO2, NiO/Al2O3 nanocomposites exhibit a satisfactory efficient photocatalytic behavior [65]. The higher methylene blue efficiency was for the NiO/TiO2 composite arising from the chitosan (NiCl2.6H2O)x//TiO2 precursor, see Table 3. This can be due to a p-n junction that can be formed acting NiO as p-NiO and TiO2 as n-TiO2. Therefore, it seems that the matrix is playing a crucial role for the NiO/TiO2 composite, and in this case, the NiO acts as the matrix rather than an active semiconductor, see Figure 22.
We have also studied the photocatalytic behavior of ReO3 prepared from the pyrolysis of the chitosan·(ReCl3)X and PSP-4-PVP·(ReCl3)X precursors [62]. The as-prepared ReO3 exhibits a moderated and high activity for ReO3 arising from chitosan and PSP-4-PVP precursors’ respectively. The composite ReO3//SiO2 prepared by solid-state pyrolysis of the chitosan·(ReCl3)X(SiO2)y and PSP-4-PVP·(ReCl3)X(SiO2)y precursors exhibit a moderate photocatalytic activity toward the degradation of methylene blue and similar to that of ReO3, see Table 4. This is the first report of the photocatalytic activity of ReO3 and ReO3//SiO2 composite.
We also tested the photocatalytic activity of IrO2 and their composite with SiO2 obtained by solid pyrolysis of the chitosan·(IrCl3)X, PSP-4-PVP·(IrCl3)X, chitosan·(IrCl3)x(SiO2)y, and PSP-4-PVP·(IrCl3)x(SiO2)y precursors.
The oxide Rh2O3 and the mixture Rh/RhO2 have band-gap values of 3 and 3.7 eV, respectively, so they could have photocatalytic activity using UV irradiation [64]. In fact, Rh2O3 and the Rh/RhO2 mixture exhibit methylene blue degradation of 70% and 78% in 300 min, respectively. To the best of our knowledge, no photodegradation of pollutants using these type Rh oxides have been reported previously.
Finally, we have measured for the first time the catalytic activity of the thoria and of their ThO2/SiO2 and ThO2/TiO2 composites. As shown in Table 5, thoria prepared from the chitosan·Th(NO3)4 precursor exhibited an activity of 66% in 300 min while that thoria prepared from PS-co-4-PVP·Th(NO3)4 precursor presents activity of 67% degradation of MB in the same time. In addition, the photocatalytic efficiency of the ThO2/SiO2 and ThO2/TiO2 composites decrease significantly, as can be viewed from Table 5. The photocatalytic activity toward methylene blue degradation follows the order ThO2 > ThO2/TiO2 > ThO2/SiO2, which can be due to the decrease in the active sites of the surface as a consequence of the encapsulation of the ThO2 into TiO2. An additional reason could be the porous morphology of ThO2, which is encapsulated inside the SiO2 and TiO2 matrices [66].

5. Probable Formation Mechanism of Nanostructures Metallic and Metal Oxides

Although the formation mechanism of nanoparticles in solution is well known [152,153,154], the studies of solid-state preparation methods are limited, and the parameters controlling both the size and the morphology of the formed nanoparticles are unknown. In this regard, we have evidenced that the pyrolysis of the {NP(OC8H12)2(OC6H4PPh2-Mn(CO)25-C5H4Me)2} precursor occurs through the intermediate formation of a layered graphite host, which is formed in the first step of the thermal solid-state reaction [16]. In addition, the formation of the nanostructured Mn2O3 and Co2O3 compounds from their respective (chitosan)(MLn)x, MLn = MnCl2, and CoCl2 macromolecular complexes precursors was confirmed to occur through an intermediate state, a layered graphitic carbon matrix, which was observed by HRTEM and Raman measurements [55]. More recently, the formation of TiO2 using several pyrolysis temperatures was confirmed with the formation of a graphite intermediate [59]. Considering all studies, a general mechanism is proposed, as shown in Figure 23.
The first step on heating involves the formation of a 3D network [15] to produce a thermally stable matrix. This step is crucial because it offset the sublimation. In our system, the first heating step could involve a cross-linking of the Chitosan and PS-co-4-PVP polymers, giving a 3D matrix containing O-M-O and H2N-M-NH2 links (for the chitosan polymer) and (pyridine)N-M-N(pyridine) bonds for the PS-co-4-PVP polymer. The following steps involve the starting of the organic carbonization, producing holes where the nanoparticles begin to nucleate. According to TG/DSC analysis, this occurs at ~400 °C for the chitosan and 360 °C for PS-co-4-PVP polymer matrices. In this intermediate stage, a layered graphitic carbon host (detected in our previous work [16]) acts as a template where the nanoparticles grow. After complete combustion, this template disappears but always remains carbon residues appearing as an ultrathin carbon shell around the nanoparticles [16].

6. Concluding Remarks

Although there are several solid-state methods to prepare nanoparticles, scarce mechanism studies have been reported, being this is a pending challenge. In this sense, control of parameters such as size and morphology of the formed nanoparticles is not known. Using the proposed solid-state method from the chitosan·MXn and PS-co-4-PVP·MXn complexes as precursors offers a reliable, general, and easy way for obtaining metal and metal oxides from all the periodic table. Using the chitosan●MLn//M′xO′y and PS-co-4-PVP●MLn//M′xO′ precursors, M/M′xO′y and MxOy/M′xO′y composites with M′xOy solid matrices can be easily obtained. However, a most detailed study of the effect of the inclusion of the metal and metal oxides nanoparticles inside the M′xOy matrix is still pending. In spite of that, several metal oxides obtained by the described solid-state method exhibit a satisfactory photocatalytic toward contaminant dyes as methylene blue. In any case, most of them need to be double-checked to obtain a general conclusion about the photocatalytic effectiveness of these oxides in order to improve them. It is also envisaged that the nanostructured metal oxides described here could significantly contribute to environmental decontamination.

Author Contributions

C.D., M.L.V. and M.Á.L.-B. conceptualized, wrote the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

Department of Chemistry, Faculty of Sciences, University of Chile; Grant PID2019-107106RB-C32 funded by MCIN/AEI/10.13039/501100011033.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hornyak, G.; Tibbals, H.F.; Dutta, J.; Moore, J. Introduction to Nanoscience and Nanotechnology; CRC Press Taylor and Francis: Boca Raton, NY, USA, 2009. [Google Scholar]
  2. Altavilla, C.; Ciliberto, E. Inorganic Nanoparticles; CRC Press Taylor and Francis: Boca Raton, NY, USA, 2011. [Google Scholar]
  3. Rao, C.N.; Muller, A.; Cheetham, A.K. The Chemistry of Nanomaterials; Wiley–VCH: Weinheim, Germany, 2004. [Google Scholar]
  4. Edelstein, A.S.; Cammarata, R.C. Nanomaterials, Synthesis and Applications; Institute of Physics Publishing: Bristol, UK, 2002. [Google Scholar]
  5. Cao, G. Nanostructures and Nanomaterial, Synthesis, Properties and Applications; Imperial College Press: London, UK, 2011. [Google Scholar]
  6. Díaz, C.; Valenzuela, M.L. Metallic Nanostructures Using Oligo and Polyphosphazenes as Template or Stabilizer in Solid State. In Encyclopedia of Nanoscience and Nanotechnology; Nalwa, H.S., Ed.; American Scientific Publishers: Valencia, CA, USA, 2010; Volume 16, pp. 239–256. [Google Scholar]
  7. Walter, E.C.; Ng, K.; Zach, M.P.; Penner, R.M.; Favier, F. Electronic Devices from electrodeposited metal nanowires. Microelectron. Eng. 2002, 61–62, 555–561. [Google Scholar] [CrossRef]
  8. Walkers, G.; Parkin, I.P. The Incorporation of noble nanoparticles into hast matrix thin films, synthesis, characterization and applications. J. Mater. Chem. 2009, 19, 574–590. [Google Scholar] [CrossRef]
  9. Scott, C.A. Epitaxial growth and properties of doped transition metal and complex oxide films. Adv. Mater. 2010, 21, 219–248. [Google Scholar]
  10. Teo, B.; Sun, X. Silicon-Based Low-Dimensional Nanomaterials and Nanodevices. Chem. Rev. 2010, 107, 1454–1532. [Google Scholar] [CrossRef] [PubMed]
  11. Khomutov, G.B.; Kislov, V.V.; Antipina, M.N.; Gainutdinov, R.V.; Gubin, S.P.; Obydenov, A.Y.; Pavlov, S.A.; Rakhnyanskaya, A.A.; Sergeev-Cherenkov, A.N.; Soldatov, E.S.; et al. Interfacial nanofabrication strategies in development of new functional nanomaterials and planar supramolecular nanostructures for nanoelectronic and nanotechnology. Microelectron. Eng. 2003, 69, 373–383. [Google Scholar] [CrossRef]
  12. Diaz, C.; Valenzuela, M.L.; Bobadilla, D. Bimetallic Au/Ag metal superstructures from macromolecular metal complexes in solid-state. J. Chil. Chem. Soc. 2013, 58, 1194–1997. [Google Scholar] [CrossRef] [Green Version]
  13. Díaz, C.; Valenzuela, M.L. Small-Molecule and High-Polymeric Phosphazenes containing oxypyridine side groups and their organometallic derivatives, Useful precursors for metal nanostructured materials. Macromolecules 2006, 39, 103–111. [Google Scholar] [CrossRef]
  14. Díaz, C.; Valenzuela, M.L. Organometallic Derivatives of Polyphosphazenes as Precursors for Metallic Nanostructured Materials. J. Inorg. Organomet. Polym. 2006, 16, 419–435. [Google Scholar] [CrossRef]
  15. Díaz, C.; Valenzuela, M.L.; Zuñiga, L.; O’Dwyer, C. Organometallic derivatives of cyclotriphosphazene as precursors of Nanostructured metallic materials, A new solid state Method. J. Inorg. Organomet. Polym. Mater. 2009, 19, 507–520. [Google Scholar] [CrossRef] [Green Version]
  16. Díaz, C.; Valenzuela, M.L.; Lavayen, V.; O’Dwyer, C. Layered Graphitic Carbon Host Formation during Liquid-free Solid State Growth of Metal Pyrophosphates. Inorg. Chem. 2012, 51, 6228–6236. [Google Scholar] [CrossRef] [Green Version]
  17. Díaz, C.; Valenzuela, M.L.; Carriedo, G.A.; Zuñiga, L.; O’Dwyer, C. Polymer/Trimer/Metal Complex Mixtures as Precursors of Gold Nanoparticles, Tuning the Morphology in the Solid-State. J. Inorg. Organomet. Polym. 2012, 22, 447–454. [Google Scholar]
  18. Díaz, C.; Valenzuela, M.L.; Cáceres, S.; O’Dwyer, C. Solution and surfactant-free growth of supported high index facet SERS active anoparticles of rhenium by phase demixing. J. Mater. Chem. A 2013, 1, 1566–1572. [Google Scholar]
  19. Díaz, C.; Valenzuela, M.L.; Cáceres, S.; O’Dwyer, C.; Diaz, R. Solvent and stabilizer free growth of Ag and Pd nanoparticles using Metallic salts/cyclotriphosphazenes mixtures. Mater. Chem. Phys. 2013, 143, 124–132. [Google Scholar] [CrossRef]
  20. Díaz, C.; Valenzuela, M.L.; Zuñiga, L.; O’Dwyer, C. Solid State Pathways to Complex Shape Evolution and Tunable Porosity During Metallic Crystal Growth. Sci. Rep. 2013, 3, 2642. [Google Scholar]
  21. Díaz, C.; Valenzuela, M.L. Organometallic-Metallic-Cyclotriphosphazene Mixtures, Solid-State Method for Metallic Nanoparticle Growth. In Nanostructures, Properties, Production Methods and Application; Nova Science Publishers: New York, NY, USA, 2013; Chapter 5. [Google Scholar]
  22. Díaz, C.; Valenzuela, M.L. A General Solid-State approach to metallic, metal oxides and Phosphates Nanoparticles. In Advances in Chemical Research; Nova Science Publishers: New York, NY, USA, 2011. [Google Scholar]
  23. Díaz, C.; Valenzuela, M.L. A general Solid-State approach to Metallic, Metal oxides and Phosphate nanoparticles. Gold Nanoparticles, Properties Synthesis and Fabrication. In Solution and Solid State Methods to Prepare Au Nanoparticles: A Comparison; Chow, P.E., Ed.; Nova Science Publishers: New York, NY, USA, 2010; Chapter 14. [Google Scholar]
  24. Larsen, T.H.; Sigman, M.; Ghezelbash, A.; Christopher-Doty, R.; Korgel, B.A. Solventless Synthesis of Copper Sulfide Nanorods by Thermolysis of a Single Source Thiolate-Derived Precursor. J. Am. Chem. Soc. 2003, 125, 5638–5639. [Google Scholar] [CrossRef] [PubMed]
  25. Sigman, M.; Ghezelbash, A.; Hanrath, T.; Saunders, A.E.; Lee, F.; Korgel, B.A. Solventless Synthesis of Monodisperse Cu2S Nanorods, Nanodisks, and Nanoplatelets. J. Am. Chem. Soc. 2003, 125, 16050–16057. [Google Scholar] [CrossRef] [PubMed]
  26. Ghezelbash, A.; Sigman, M.; Korgel, B.A. Solventless Synthesis of Nickel Sulfide Nanorods and Triangular Nanoprisms. Nano Lett. 2004, 4, 537–542. [Google Scholar] [CrossRef]
  27. Sigman, M.; Korgel, B.A. Solventless Synthesis of Bi2S3 (Bismuthinite) Nanorods, Nanowires, and Nanofabric. Chem. Mater. 2005, 17, 1655–1660. [Google Scholar] [CrossRef]
  28. Han, Y.C.; Cha, H.G.; Kim, C.h.W.; Kim, Y.H.; Kang, Y.S. Synthesis of Highly Magnetized Iron Nanoparticles by a Solventless Thermal Decomposition Method. J. Phys. Chem. C 2007, 111, 6275–6280. [Google Scholar] [CrossRef]
  29. Farhadi, S.; Roostaei-Zaniyani, Z. Preparation and characterization of NiO nanoparticles from thermal decomposition of the Ni(en)3(NO3)2 complex, A facile and low-temperature route. Polyhedron 2011, 30, 971–975. [Google Scholar] [CrossRef]
  30. Li, X.; Zhang, X.; Li, Z.; Qian, Y. Synthesis and characteristics of NiO nanoparticles by thermal decomposition of nickel dimethylglyoximate rods. Solid State Commun. 2006, 137, 581–584. [Google Scholar] [CrossRef]
  31. Davar, F.; Salavati-Niasari, M.; Mir, N.; Saberyan, K.; Monemzadeh, M. Thermal decomposition route for synthesis of Mn3O4 nanoparticles in presence of a novel precursor. Polyedron 2010, 29, 1747–1753. [Google Scholar] [CrossRef]
  32. Yanh, Z.; Zhang, Y.; Zhang, W.; Wang, X.; Qian, Y.; Weng, X.; Yang, S. Nanorods of manganese oxides, Synthesis, characterization and catalytic application. J. Solid State Chem. 2006, 179, 679–684. [Google Scholar]
  33. Soofivand, F.; Salavati-Niasari, M.; Mohandes, F. Novel precursor-assisted synthesis and characterization of zinc oxide nanoparticles/nanofibers. Mater. Lett. 2013, 98, 55–58. [Google Scholar] [CrossRef]
  34. Farhadi, S.; Pouzare, K.; Sadedhinejad, S. Simple preparation of ferromagnetic CO3O4 nanoparticles by thermal dissociation of the CoII(NH3)6(NO3)2 complex at low termperature. J. NanoStruct. Chem. 2013, 3, 16. [Google Scholar] [CrossRef] [Green Version]
  35. Randhawa, B.S.; Gandotra, K. A comparative study on the thermal decomposition of some transition metal carboxylates. J. Therm. Anal. Calorim. 2006, 85, 417–424. [Google Scholar] [CrossRef]
  36. Chunxiang Li Weng, K.L. Thermolysis of Polymeric Ru(CO)4.n to Metallic Ruthenium, Molecular Shape of the Precursor Affects the Nanoparticle Shape. Langmuir 2008, 24, 12040–12041. [Google Scholar]
  37. Nelson, J.M.; Nguyen, P.; Petersen, R.; Rengel, H.; Macdonald, P.M.; Lough, I.J.; Manners, I.; Raju, N.P.; Greedan, J.E.; Barlow, S.; et al. Thermal Ring-Opening Polymerization of Hydrocarbon-Bridged 2.Ferrocenophanes, Synthesis and Properties of Poly(ferrocenylethy1ene)s and Their Charge-Transfer Polymer Salts with Tetracyanoethylene. Chem. Eur. J. 1997, 3, 573–584. [Google Scholar] [CrossRef]
  38. Tang, B.Z.; Petersen, R.; Foucher, D.A.; Lough, A.; Coombs, N.; Sodhi, R. Novel Ceramic and Organometallic Depolymerization Products from PoIy(ferrocenyIsiIanes) via PyroIysis. J. Chem. Soc. Chem. Commun. 1993, 523–525. [Google Scholar] [CrossRef] [Green Version]
  39. Chang, B.S.; Brijith, T.; Chen, J.; Tevis, I.D.; Karanja, P.; Çınar, S.; Venkatesh, A.; Rossini, A.J.; Thuo, M.M. Ambient synthesis of nanomaterials by in situ heterogeneous metal/ligand reactions. Nanoscale 2019, 11, 14060–14069. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Li, C.; Zhong, Z.; Leong, W.K. Organometallic Clusters As Precursors for Metallic Nanoparticles, Effect of Cluster Size, Ligand Set, and Decomposition Method. Langmuir 2008, 24, 10427–10431. [Google Scholar] [CrossRef] [PubMed]
  41. Cai, B.; Akkiraju, K.; Mounfield, W.P.; Wang, Z.; Li, X.; Huang, B.; Yuan, S.; Su, D.; Roman-Leshkov, Y.; Shao-Horn, Y. Solid-State Gelation for Nanostructured Perovskite Oxide Aerogel. Chem. Mater. 2019, 31, 9422–9429. [Google Scholar] [CrossRef]
  42. Wang, B.; Zhang, C.; Zheng, W.; Zhang, Q.; Bao, Z.; Kong, L.; Li, L. Large-Scale Synthesis of Highly Luminescent Perovskite Nanocrystals by Template-Assisted Solid-State Reaction at 800. Chem. Mater. 2020, 32, 308–314. [Google Scholar] [CrossRef]
  43. Mao, Y.; Banerjee, S.; Wong, S.S. Large-Scale Synthesis of Single-Crystalline Perovskite Nanostructures. J. Am. Chem. Soc. 2003, 125, 15718–15719. [Google Scholar] [CrossRef]
  44. Schmitt, W.; Hill, J.P.; Malik, S.; Volkert, C.A.; Ichinose, I.; Anson, C.; Powell, A.K. Thermolysis of a Hybrid Organic–Inorganic Supramolecular Coordination Assembly, Templating the Formation of Nanostructured Fibrous Materials and Carbon-Based Microcapsules. Angew. Chem. Int. Ed. 2005, 44, 7048–7053. [Google Scholar] [CrossRef] [PubMed]
  45. Reda, G.M.; Fan, H.; Tian, H. Room-temperature solid state synthesis of Co3O4/ZnO p–n heterostructure and its photocatalytic activity. Adv. Powder Technol. 2017, 28, 953–963. [Google Scholar] [CrossRef]
  46. Nalluri, S.R.; Nagarjuna, R.; Patra, D.; Ganesan, R.; Balaj, G. Large Scale Solid-state Synthesis of Catalytically Active Fe3O4@M (M =Au, Ag and Au-Ag alloy) Core-shell Nanostructures. Sci. Rep. 2019, 9, 6603. [Google Scholar] [CrossRef]
  47. Dey, A.; Zubko, M.; Kusz, J.; Reddy, V.R.; Banerjee, A.; Bhattacharjee, A. Thermal synthesis of Hematite nanoparticles, Structural, magnetic and morphological characterizations. Int. J. Nano Dimens. 2020, 11, 188–198. [Google Scholar]
  48. Das, R.; Pachfule, P.; Banerjee, R.; Poddar, P. Metal and Metal oxide nanoparticle synthesis from metal organic frameworks (MOFs), finding the border of metal and metal oxides. Nanoscale 2012, 4, 591–599. [Google Scholar] [CrossRef] [PubMed]
  49. Rong, K.; Wei, J.; Huang, L.; Fang, Y.; Dong, S. Synthesis of low dimensional hierarchical transition metal oxides via a direct deep eutectic solvent calcining method for enhanced oxygen evolution catalysis. Nanoscale 2020, 12, 20719–20725. [Google Scholar] [CrossRef]
  50. Zhu, J.; Jiang, Y.; Lu, Z.; Zhao Ch Xie, L.; Chen, L.; Duan, J. Single-crystal Cr2O3 nanoplates with differing crystalinities, derived from trinuclear complexes and embedded in a carbon matrix, as an electrode material for supercapacitors. J. Colloid Interface Sci. 2017, 498, 351–363. [Google Scholar] [CrossRef] [PubMed]
  51. Yuan, Y.; Chen, L.; Yang, R.; Lu, X.; Peng, H.; Luo, Z. Solid-state synthesis and characterization of core–shell CoFe2O4–carbon composite nanoparticles from a heterometallic trinuclear complex. Mater. Lett. 2012, 71, 123–126. [Google Scholar] [CrossRef]
  52. Díaz, C.; Valenzuela, M.L.; Lavayen, V.; Mendoza, K.; Peña, O.; O’Dwyer, C. Nanostructured copper oxides and phosphates from a new solid-state route. Inorg. Chim. Acta 2011, 377, 5–11. [Google Scholar] [CrossRef]
  53. Díaz, C.; Platoni, S.; Molina, A.; Valenzuela, M.L.; Geaney, H.; O’Dwyer, C. Novel Solid-State Route to Nanostructured Tin, Zinc and Cerium Oxides as Potential Materials for Sensors. J. Nanosci. Nanotechnol. 2014, 14, 7648–7653. [Google Scholar] [CrossRef]
  54. Diaz, C.; Barrera, G.; Segovia, M.; Valenzuela, M.L.; Osiak, M.; O’Dwyer, C. Solvent-less method for efficient photocatalytic α-Fe2O3 nanoparticles for using macromolecular polymeric precursors. New J. Chem. 2016, 40, 6768–6776. [Google Scholar] [CrossRef]
  55. Diaz, C.; Valenzuela, M.L.; Laguna, M.A.; Orera, A.; Bobadilla, D.; Abarca, S.; Peña, O. Synthesis and Magnetic Properties of Nanostructured metallic Co, Mn and Ni oxide materials obtained from solid-state macromolecular complex precursors. RSC Adv. 2017, 7, 27729–27736. [Google Scholar] [CrossRef] [Green Version]
  56. Diaz, C.; Valenzuela, M.L.; Bobadilla, D.; Laguna-Bercero, M.A. Bimetallic Au//Ag Alloys Inside SiO2 using a solid-state method. J. Clust. Chem. 2017, 28, 2809–2815. [Google Scholar] [CrossRef] [Green Version]
  57. Diaz, C.; Valenzuela, M.L.; Garcia, C.; De la Campa, R.; Soto, A.-P. Solid-state synthesis of pure and doped lanthanides oxide nanomaterials by using polymer templates. Study of their luminescent properties. Mater. Lett. 2017, 209, 111–114. [Google Scholar] [CrossRef]
  58. Diaz, C.; Valenzuela, M.L.; Segovia, M.; De la Campa, R.; Soto, A.-P. Solution, Solid-State Two Step Synthesis and Optical Properties of ZnO and SnO Nanoparticles and Their Nanocomposites with SiO2. J. Clust. Sci. 2018, 29, 251–266. [Google Scholar] [CrossRef] [Green Version]
  59. Allende, P.; Laguna, M.A.; Barrientos, L.; Valenzuela, M.L.; Diaz, C. Solid State tuning Morphology, Crystal Phase and Size through Metal Macromolecular Complexes and Its Significance in the Photocatalytic Response. ACS Appl. Energy Mater. 2018, 1, 3159–3170. [Google Scholar] [CrossRef]
  60. Diaz, C.; Carrillo, D.; De la Campa, R.; Soto, A.-P.; Valenzuela, M.L. Solid-State synthesis of LnOCl/Ln2O3 (Ln = Eu, Nd) by using chitosan and PS-co-P4VP as polymeric supports. J. Rare Earth 2018, 36, 1326–1332. [Google Scholar] [CrossRef]
  61. Allende, P.; Barrientos, L.; Orera, A.; Laguna-Bercero, M.A.; Salazar, N.; Valenzuela, M.L.; Diaz, C. TiO2/SiO2 Composite for Efficient Protection of UVA and UVB Rays Through of a Solvent-Less Synthesis. J. Clust. Sci. 2019, 30, 1511–1517. [Google Scholar] [CrossRef]
  62. Diaz, C.; Valenzuela, M.L.; Cifuentes-Vaca, O.; Segovia, M.; Laguna-Bercero, M.A. Incorporation of Nanostructured ReO3 in Silica Matrix and Their Activity Toward Photodegradation of Blue Methylene. J. Inorg. Organomet. Polym. Mater. 2020, 30, 1726–1734. [Google Scholar] [CrossRef]
  63. Diaz, C.; Valenzuela, M.L.; Cifuentes-Vaca, O.; Segovia, M.; Laguna-Bercero, M.A. Iridium nanostructured metal oxide, its inclusion in Silica matrix and their activity toward Photodegradation of Methylene Blue. Mater. Chem. Phys. 2020, 252, 123276–123286. [Google Scholar] [CrossRef]
  64. Diaz, C.; Valenzuela, M.L.; Cifuentes-Vaca, O.; Segovia, M. Polymer precursors effect in the macromolecular metal-polymer on the Rh/RhO2/Rh2O3 phase using solvent-less synthesis and its photocatalytic activity. J. Inorg. Organomet. Polym. Mater. 2020, 30, 4702–4708. [Google Scholar] [CrossRef]
  65. Diaz, C.; Valenzuela, M.L.; Cifuentes-Vaca, O.; Segovia, M.; Laguna-Bercero, M.A. Incorporation of NiO into SiO2, TiO2, Al2O3, and Na4.2Ca2.8(Si6O18) Matrices, Medium Effect on the Optical Properties and Catalytic Degradation of Methylene Blue. Nanomaterials 2020, 10, 2470. [Google Scholar] [CrossRef]
  66. Diaz, C.; Valenzuela, M.L.; Laguna-Bercero, M.A.; Mendoza, K.; Cartes, P. Solventless preparation of thoria, their inclusion inside SiO2 and TiO2, their luminiscent properties and their photocataltytic behavior. ACS Omega 2021, 6, 9391–9400. [Google Scholar] [CrossRef] [PubMed]
  67. Chen, X.; Mao, S.S. Titanium Dioxide Nanomaterials, Synthesis, Properties, Modifications, and Applications. Chem. Rev. 2007, 107, 2891–2959. [Google Scholar] [CrossRef] [PubMed]
  68. Ismail, A.A.; Bahnemannb, D.W. Mesoporous titania photocatalysts, preparation, characterization and reaction mechanisms. J. Mater. Chem. 2011, 21, 11686–11707. [Google Scholar] [CrossRef] [Green Version]
  69. Wang, Y.; Cao, G. Synthesis and Enhanced Intercalation Properties of Nanostructured Vanadium Oxides. Chem. Mater. 2006, 18, 2787–2804. [Google Scholar] [CrossRef]
  70. Mao, C.J.; Pan, H.C.; Wu, X.C.; Zhu, J.J.; Chen, H.Y. Sonochemical Route for Self-Assembled V2O5 Bundles with Spindle-like Mosphology and their Novel Application in Serum Albumin Sensing. J. Phys. Chem. 2006, 110, 14709–14713. [Google Scholar]
  71. Avansi, W.; Ribeiro, C.; Leite, E.; Mastelaro, V. Vanadium Pentoxide Nanostructured, An Effective Control of Morphology and Crystal Strucutred in Hydrothermal Conditions. Cryst. Growth Des. 2009, 9, 3626–3631. [Google Scholar] [CrossRef]
  72. Diaz, C.; Barrera, G.; Segovia, M.; Valenzuela, M.L.; Osiak, M.; O’Dwyer, C. Crystallizing Vanadium Pentoxide Nanostructures in the Solid State using Modified Block co-Polymer and Chitosan Complexes. J. Nanomater. 2015, 2015, 105157. [Google Scholar] [CrossRef] [Green Version]
  73. Fei, H.L.; Liu, M.; Zhou, H.J.; Sun, P.C.; Ding, D.T.; Chen, T.H. Synthesis of V2O5 micro-architectures via in situ generation of single-crystalline nanoparticles. Solid State Sci. 2009, 11, 102–107. [Google Scholar] [CrossRef]
  74. Díaz, C.; Valenzuela, M.L.; Yutronic, N.; Villalobos, V.; Barrera, G. Nanostructured VOx/VO(PO4)n Using Solid-State Vanadium Containing Phosphazene Precursors, A Uselful Potential Bi-Catalyst System. J. Clust Sci. 2011, 22, 693–704. [Google Scholar] [CrossRef]
  75. Zhou, Y.; Qiu, Z.; Lu, M.; Zhang, A.; Ma, Q. Preparation and characterization of V2O5 macro-plates. Mater. Lett. 2007, 61, 4073–4075. [Google Scholar] [CrossRef]
  76. Diaz, C.; Valenzuela, M.L.; Zepeda, L.; Herrera, P.; Valenzuela, C. General group VI transition nanostructured metal oxides and their inclusion into solid matrices by a solution-solid approach. J. Chil. Chem. Soc. 2021, 66. in press. [Google Scholar]
  77. Han, Y.; Chen, F.; Zhong, Z.; Ramesh, K.; Chen, L.; Widjaja, E. Controlled Synthesis, Characterization, and Catalytic Properties of Mn2O3 and Mn3O4 Nanoparticles Supported on Mesoporous Silica SBA-15. J. Phys. Chem. B 2006, 110, 24450–24456. [Google Scholar] [CrossRef]
  78. Laurent, S.; Forge, D.; Port, M.; Roch, A.; Robic, C.; Vander Elst, L.; Muller, R. Magnetic Iron Oxide Nanoparticles, Synthesis, Stabilization, Vectorization, Physicochemical Characterizations and Biological Applications. Chem. Rev. 2008, 108, 2064–2110. [Google Scholar] [CrossRef]
  79. Teja, A.; Koh, P.Y. Synthesis propieties and applications of magnetic iron oxide nanoparticles. Prog. Cryst. Growth Charact. Mater. 2009, 55, 22–45. [Google Scholar] [CrossRef]
  80. Patra, A.K.; Kundu, S.K.; Bhaumik, A.; Kim, D. Morphology evolution of single-crystalline hematite nanocrystals: Magnetically recoverable nanocatalysts for enhanced facet-driven photoredox activity. Nanoscale 2016, 8, 365. [Google Scholar] [CrossRef] [PubMed]
  81. Zheng, Y.; Cheng, Y.; Wang, Y.; Bao, F.; Zhou, L.; Wei, X.; Zhang, Y.; Zheng, Q. Quasicubic α-Fe2O3 Nanoparticles with Excellent Catalytic Performance. J. Phys. Chem. 2006, 110, 3093–3097. [Google Scholar] [CrossRef]
  82. Hua, J.; Gengsheng, J. Hydrothermal synthesis abd characterization of monodispere α-Fe2O3 Nanoparticles. Mater. Lett. 2009, 63, 2725–2727. [Google Scholar] [CrossRef]
  83. Liu, Y.; Yu, C.; Dai, W.; Gao, X.; Qian, H.; Hu, Y.; Hu, X. One-post solvothermal synthesis of multi-shelled α-Fe2O3 hollow spheres with enhaced visible-light photocatalytic activity. J. Alloy. Compd. 2013, 551, 440–443. [Google Scholar] [CrossRef]
  84. Jiu, J.; Ge, Y.; Li, X.; Nie, L. Preparation of Co3O4 nanoparticles by a polymer conbution route. Mater. Lett. 2002, 54, 260–263. [Google Scholar] [CrossRef]
  85. Li, Y.; Tan, B.; Wu, Y. Mesoporous Co3O4 Nanowires Arrays for Lithium Ion Batteries with High Capacity and rate Capability. Nano Lett. 2008, 8, 265–270. [Google Scholar] [CrossRef]
  86. Deng, J.; Kang, L.; Bai, G.; Li, Y.; Li, P.; Liu, X.; Yang, Y.; Gao, F.; Liang, W. Solution combustion synthesis of cobalt oxides (Co3O4 and Co3O4/CoO) nanoparticles as supercapacitor electrode materials. Electrochim. Acta 2014, 132, 127–135. [Google Scholar] [CrossRef]
  87. Salvati-Niasari, M.; Khansari, A.; Davar, F. Synthesis and characterization of cobalt oxides nanoparticles by thermal process. Inorg. Chem. Acta 2009, 362, 4937–4942. [Google Scholar] [CrossRef]
  88. Nassar, M.Y.; Ahmed, I.S. Hydrothermal synthesis of cobalt carbonates using different counter ions, An efficient precursor to nano-sized cobalt oxide (Co3O4). Polyhedron 2011, 30, 2431–2437. [Google Scholar] [CrossRef]
  89. Palacios-Hernandez, P.; Hirata-Flores, G.; Contreras-Lopez, O.; Mendoza-Sanchez, M.; Valeriano-Arreola, I.; Gonzalez-Vergara, E.; Mendez-Rojas, M. Synthesis of Cu and Co metal oxide nanoparticles from thermal decomposition of tastrate complexes. Inorg. Chem. Acta 2012, 392, 277–282. [Google Scholar] [CrossRef]
  90. Kalam, A.; Al-Sehemi, A.; Al-Shihri, A.; Du, G.; Tokeer, A. Synthesis and Characterization if NiO nanoparticles by thermal decomposition of nickel linoleate and their optical properties. Mater. Charact. 2012, 68, 77–81. [Google Scholar] [CrossRef]
  91. Wang, C.; Li, J.; Liang, X.; Zhang, Y.; Guo, G. Photocatalytic organic pollutants degradation in metal–organic frameworks. Energy Environ. Sci. 2014, 7, 2831–2867. [Google Scholar] [CrossRef]
  92. Ukoba, K.O.; Eloka-Eboca, A.C.; Inambao, F.L. Review of nanostructured NiO thin film deposition using the spray pyrolysis technique. Renew. Sustain. Energy Rev. 2018, 82, 2900–2915. [Google Scholar] [CrossRef]
  93. Rowashden-Omary, M.; Lopez-Luzuriaga, J.M.; Rashdan, M.; Elbjeirami, O.; Monge, M.; Rodriguez-Castillo, M.; Laguna, A. Golden Metallopolymers with an Active T State via Coordination of Poly(4-vinyl)pyridine to Pentahalophenyl-Gold (I) Precursors. J. Am. Chem. Soc. 2009, 131, 3824–3825. [Google Scholar] [CrossRef]
  94. Diaz, C.; Valenzuela, M.L.; Soto, K.; Laguna-Bercero, M.A. Incorporation of Au and Ag Nanostructures inside SiO2. J. Chilean Chem. Soc. 2019, 64, 4502–4506. [Google Scholar] [CrossRef] [Green Version]
  95. Tappan, B.; Steiner, S.; Luther, E. Nonporous Metal Foams. Angew. Chem. Int. Ed. 2010, 49, 4544–4565. [Google Scholar] [CrossRef]
  96. Yaqoob, A.A.; Umar, K.; Nasir, M.; Ibrahim, M. Silver nanoparticles, various methods of synthesis, size afecting factors and their potential applications—A review. Appl. Nanosci. 2020, 10, 1369–1378. [Google Scholar] [CrossRef]
  97. Syafiuddin, A.; Salmiati Salim, M.R.; Hong Kueh, A.B.; Hadibarata, T.; Nure, H.A. Review of Silver Nanoparticles, Research Trends, Global Consumption, Synthesis, Properties, and Future Challenges. J. Chin. Chem. Soc. 2017, 64, 732–756. [Google Scholar] [CrossRef]
  98. Chen, A.; Holt-Hindle, P. Platinum-Based Nanostructured Materials, Synthesis, Properties and Applications. Chem. Rev. 2010, 110, 3767–3804. [Google Scholar] [CrossRef]
  99. Diaz, C.; Valenzuela, M.L.; Baez, R.; Segovia, M. Solid State Morphology and Size Tuning of Nanostructured Platinum Using Macromolecular Complexes. J. Chil. Chem. Soc. 2015, 60, 2986–2990. [Google Scholar] [CrossRef] [Green Version]
  100. Wu, J.; Qi, L.; You, H.; Gross, A.; Li, J.; Yang, H. Icosahedral Platinum Alloy Nancrystals with Enhanced Electrocatalytic activities. J. Am. Chem. Soc. 2012, 134, 11880–11883. [Google Scholar] [CrossRef]
  101. Leong, G.J.; Schulze, M.C.; Strand, M.B.; Maloney, D.; Frisco, S.L.; Dinh, H.N.; Pivovar, B.; Richards, R.M. Shape-directed platinum nanoparticle synthesis, nanoscale design of novel catalysts. Appl. Organomet. Chem. 2014, 28, 1–17. [Google Scholar] [CrossRef]
  102. Peng, H.; Yang, H. Designer platinum nanoparticles, Control of shape, composition in alloy, nanostrucutred and electrocatalytic property. Nano Today 2009, 4, 143–164. [Google Scholar] [CrossRef]
  103. Chen, J.; Lim, B.; Lee, E.P.; Xia, Y. Shape-controlled synthesis of platinum nanocrystals for catalytic and electrocatalytic applications. Nano Today 2009, 4, 81–95. [Google Scholar] [CrossRef]
  104. Cotton, F.A.; Wilkinson, G. Chapter 22 and 30. In Advanced Inorganic Chemistry; John Wiley and Sons: New York, NY, USA, 1980. [Google Scholar]
  105. Jin, R. The impacts of nanotechnology on catalysis by precious metal nanoparticles. Nanotechnol. Rev. 2012, 1, 31–56. [Google Scholar] [CrossRef]
  106. Liu, L.; Corma, A. Metal Catalysts for Heterogeneous Catalysis: From Single Atoms to Nanoclusters and Nanoparticles. Chem. Rev. 2018, 118, 4981–5079. [Google Scholar] [CrossRef] [Green Version]
  107. Chen, R.S.; Korotcov, A.; Huiang, A.S.; Tsai, D. One-dimensional conductive IrO2 nanocrystals. Nanotechnology 2006, 17, 67–87. [Google Scholar] [CrossRef] [Green Version]
  108. Woo, H.; Shim, H.S.; Myung, J.H.; Lee, C. Annealing effect on the structural properties of IrO2. Vacuum 2008, 82, 1400–1403. [Google Scholar]
  109. Ortel, E.; Reier, T.; Strasser, P.; Kraehnert, R. Mesoporous film template by PEO-PB-PEO block-copolymers, Self-Assembly, Cristalización behaviour and electrocatalytic performance. Chem. Mater. 2011, 23, 3201–3209. [Google Scholar] [CrossRef]
  110. Zhao, Y.; Hernandez, E.A.; Vargas-Barbosa, N.M.; Dysart, J.L.; Mallouk, T.E. A high yield Synthesis of ligand-free iridium oxide nanoparticles with high electrocatalytic activity. J. Phys. Chem. Lett. 2011, 2, 402–406. [Google Scholar] [CrossRef]
  111. Brewer, D.; Wicajksana, J.; Maria, A.; Kingon, S. Franzen, Investigation of the electrical and optical properties of iridium oxide by reflectance FTIR spectroscopy and density functional theory calculations. Chem. Phys. 2005, 313, 25–31. [Google Scholar] [CrossRef]
  112. Xu, D.; Diao, P.; Jin, T.; Wu, Q.; Liu, X.; Guo, X.; Gong, H.; Li, F.; Xiang, M.; Ronghai, Y. Iridium oxide nanoparticles and Iridium/Iridium Oxide Nanocompósites, Photochemical Fabrication and Application in Catalytic Rediction of Notrophenol. ACS Appl. Interfaces 2015, 7, 16738–16749. [Google Scholar] [CrossRef] [PubMed]
  113. Biswas, K.; Rao, C.N. Synthesis and characterization of Nano crystals of oxide metals RuO2, IrO2, and ReO3. J. Nanosci. Nanotechnol. 2007, 7, 1969–1974. [Google Scholar] [CrossRef] [PubMed]
  114. Lee, Y.M.; Suntivich, J.; May, K.J.; Perry, E.E.; Shao, H. Synthesis and activities of Rutile IrO2 and RuO2 nanoparticles for oxygen evolution in acid alkaline solutions. J. Phys. Chem. Lett. 2011, 2, 402–406. [Google Scholar] [CrossRef]
  115. Quinson, J. Surfactant-Free Precious Metal Colloidal Nanoparticles for Catalysis. Front. Nanotechnol. 2021, 3, 770281. [Google Scholar] [CrossRef]
  116. Fernandez-Garcia, M.; Martinez-Arias, A.; Hanson, J.; Rodriguez, C. Nanostructured Oxides in Chemistry, Characterization and Properties. J. Am. Chem. Rev. 2004, 104, 4063–4104. [Google Scholar] [CrossRef]
  117. Kim, Y.L.; Ha, Y.; Lee, N.S.; Kim, J.G.; Baik, J.M.; Le, C.; Yoon, K.; Lee, Y.; Kim, M.H. Hybrid architecture of rhodium oxide nanofibers and ruthenium oxide nanowires for electrocatalysts. J. Alloy. Compd. 2016, 663, 574–580. [Google Scholar] [CrossRef]
  118. Bai, J.; Han, S.-H.; Peng, R.; Zheng, J.-H.; Jiang, J.-X.; Chen, Y. Ultrathin Rhodium Oxide Nanosheet Nanoassemblies, Synthesis, Morphological Stability, and Electrocatalytic Application. ACS. Appl. Mater Interfaces 2017, 9, 17195–17200. [Google Scholar] [CrossRef]
  119. Shimura, K.; Kawai, H.; Yoshida, T.; Yoshida, H. Simultaneously photodeposited rhodium metal and oxide nanoparticles promoting photocatalytic hydrogen production. Chem. Commun. 2011, 47, 8958–8960. [Google Scholar] [CrossRef] [PubMed]
  120. Saric, A.; Popovic, S.; Music, S. Formation of crystalline phases by thermal treatment of amorphous rhodium hydrous oxide. Mater. Lett. 2002, 55, 145–151. [Google Scholar] [CrossRef]
  121. Saric, A.; Popovic, S.; Trojko, R.; Music, S. The thermal behavior of amorphous rhodium hydrous oxide. J. Alloys Compd. 2011, 320, 140–148. [Google Scholar] [CrossRef]
  122. Kibis, L.S.; Stadnichenko, A.I.; Koscheev, S.V.; Zaikovskii, V. Boron in A.I. XPS Study of Nanostructured Rhodium Oxide Film Comprising Rh4+ Species. J. Phys. Chem. 2016, 120, 1942–19150. [Google Scholar]
  123. Tricoli, A.; Righettoni, M.; Dupont, L.; Teleki, A. Semiconductor Gas Sensors, Dry Synthesis and Application. Angew. Chem. Int. Ed. 2010, 49, 7632–7659. [Google Scholar] [CrossRef]
  124. Wang, H.; Rogach, A.L. Hierarchical SnO2 Nanostructures, Recent Advances in Desing, Synthesis and Applications. Chem. Mater. 2004, 26, 123–133. [Google Scholar] [CrossRef]
  125. Ahmad, M.; Zhu, J. ZnO based advanced functional nanostructures, synthesis, properties and applications. J. Mater. Chem. 2011, 21, 599–614. [Google Scholar] [CrossRef]
  126. Li, H.; Wang, X.; Huang, D.; Chen, G. Recent advances of lanthanide-doped upconversion nanoparticles for biological applications. Nanotechnology 2020, 31, 072001. [Google Scholar] [CrossRef]
  127. Qin, X.; Xu, J.; Wu, Y.; Liu, X. Energy-Transfer Editing in Lanthanide-Activated Upconversion Nanocrystals: A Toolbox for Emerging Applications. ACS Cent. Sci. 2019, 5, 29–42. [Google Scholar] [CrossRef]
  128. Binnemans, K. Lanthanide-Based Luminescent Hybrid Materials. Chem. Rev. 2009, 109, 4283–4374. [Google Scholar] [CrossRef] [Green Version]
  129. Bunzli, J.C. Benefiting from the Unique Properties of Lanthanide Ions. Acc Chem Res. 2006, 39, 53–61. [Google Scholar] [CrossRef] [PubMed]
  130. Dong, H.; Sun, L.D.; Yan, C.H. Basic understanding of the lanthanide related up conversion emissions. Nanoscale 2013, 5, 5703–5714. [Google Scholar] [CrossRef] [PubMed]
  131. Si, R.; Zhang, Y.W.; You, L.P.; Yan, C. Rare-Earth Nanopolyhedra, Nanoplates and Nanodisks. Angew. Chem. Int. Ed. 2005, 44, 3256–3260. [Google Scholar] [CrossRef] [PubMed]
  132. Kort, K.R.; Banerjee, S. Shape-Controlled Synthesis of Well-Defined Matlockite LnOVl (Ln, La, Ce, Gd, Dy) Nanocrystals by a Novel Non-Hydrolytic Approach. Inorg. Chem. 2011, 50, 5539–5544. [Google Scholar] [CrossRef]
  133. Du, Y.P.; Zhang, Y.W.; Sun, L.D.; Yan, C.H. Atomically Efficient Synthesis of Self Assembled Monodisperse and Ultrathin Lanthanide Oxychloride Nanoplates. J. Am. Chem. Soc. 2009, 131, 3162–3163. [Google Scholar] [CrossRef] [PubMed]
  134. Cotton, A.; Wilkinson, G.; Murillo, C.A.; Bochmann, M. Chapter 20. In Advanced Inorganic Chemistry; John Wiley and Sons: New York, NY, USA, 1999. [Google Scholar]
  135. Lin, Z.W.; Kuang, Q.; Lian, W.; Jiang, Z.Y.; Xie, Z.X.; Huang, R.B.; Zheng, L.S. Preparation and Optical Properties of ThO2 and Eu-Doped ThO2 Nanotubes by the Sol-Gel Method Combined with Porous Anodic Aluminum Oxide Template. J. Phys. Chem. B 2006, 110, 23007–23011. [Google Scholar] [CrossRef]
  136. Hudry, D.; Apostolidis, C.; Walter, O.; Gouder, T.; Courtois, E.; Kubel, C.; Meyer, D. Non-aqueous Synthesis of Isotropic and Anisotropic Actinide Oxide Nanocrystals. Chem. Eur. J. 2012, 18, 8283–8287. [Google Scholar] [CrossRef]
  137. Hudry, D.; Apostolidis, C.; Walter, O.; Gouder, H.; Courtois, E.; Kubel, C.; Meyer, D. Controlled Synthesis of Thorium and Uranium Oxide Nanocrystals. Chem. Eur. J. 2013, 19, 5297–5305. [Google Scholar] [CrossRef]
  138. Tripathi, V.K.; Narajan, R. Sol–Gel Synthesis of High-Purity Actinide Oxide ThO2 and Its Solid Solutions with Technologically Important Tin and Zinc Ions. Inorg. Chem. 2016, 55, 12798–12806. [Google Scholar] [CrossRef]
  139. Armelao, L.; Barreca, D.; Bottaro, G.; Gasparotto, A.; Gross, S.; Maragnob, C.; Tondello, E. Recent trends on nanocomposites based on Cu, Ag and Au clusters: A closer look. Coord. Chem. Rev. 2006, 250, 1294–1314. [Google Scholar] [CrossRef]
  140. Liu, S.; Han, M.Y. Silica-Coated Metal Nanoparticles. Chem. Asian J. 2010, 5, 36–45. [Google Scholar] [CrossRef] [PubMed]
  141. Zhang, Y.; Mei, J.; Yan, C.; Liao, T.; Bell, J.; Sun, Z. Bioinspired 2D Nanomaterials for Sustainable Applications. Adv. Mater. 2019, 32, 1902806. [Google Scholar] [CrossRef]
  142. Adeyemo, A.A.; Adeoye, I.O.; Bello, O.S. Metal organic frameworks as adsorbents for dye adsorption, overview, prospects and future challenges Toxicol. Environ. Chem. 2012, 94, 1845–1863. [Google Scholar]
  143. Ali, I.; Asim, M.; Khan, T.A. Low cost adsorbents for the removal of organic pollutants from wastewater. J. Environ. Manage. 2012, 113, 267–276. [Google Scholar] [CrossRef] [PubMed]
  144. Padmanabhan, O.P.; Sreekumar, K.; Sengupta, P.; Dey, G.; Werrier, K. Nano-crystalline titanium dioxide formed by reactive plasma synthesis. Vacuum 2006, 80, 1252–1255. [Google Scholar] [CrossRef]
  145. Gaya, U.I.; Abdullah, A.H. Heterogeneous photocatalytic degradation of organic contaminants over titanium dioxide, A review of fundamentals, progress and problems. J. Photochem. Photobiol. C 2008, 9, 1–12. [Google Scholar] [CrossRef]
  146. Chong, M.N.; Jin, B.; Chow, C.W.; Saint, C. Recent developments in photocatalytic water treatment technology: A review. Water Res. 2010, 44, 2997–3027. [Google Scholar] [CrossRef]
  147. Ray, C.; Pai, T. Recent advances of metal–metal oxide nanocomposites and their tailored nanostructures in numerous catalytic applications. J. Mater. Chem. 2017, 5, 9465–9478. [Google Scholar] [CrossRef]
  148. Preeti, S.; Abdullah, M.M.; Saiga, I. Role of Nanomaterials and Their Applications as Photo-catalst and Sensor: A review. Nano Res. Appl. 2016, 2, 1–10. [Google Scholar]
  149. Nikam, A.V.; Prasad, B.L.V.; Kulkarnia, A.A. Wet Chemical Synthesis of Metal Oxide Nanoparticles, A Review. Cryst. Eng. Comm. 2018, 20, 5091–5107. [Google Scholar] [CrossRef]
  150. Oska, G. Metal oxide nanoparticles, synthesis, characterization and application. J. Sol.-Gel. Sci. Techn. 2006, 37, 161–164. [Google Scholar] [CrossRef]
  151. Guo, H.; Li, H.; Fernandez, D.; Willis, S.; Jarvis, K.; Henkelman, G.R.; Humphrey, S.M. Stabilizer-Free CuIr Alloy Nanoparticle Catalysts. Chem. Mater. 2019, 31, 10225–10235. [Google Scholar] [CrossRef]
  152. Mozaffari, S.; Li, W.; Thompson, C.; Ivanov, S.; Seifert, S.; Lee, B.; Kovarik, L.; Karim, A. Colloidal nanoparticle size control, experimental and kinetic modeling investigation of the ligand–metal binding role in controlling the nucleation and growth kinetics. Nanoscale 2017, 9, 13772–13785. [Google Scholar] [CrossRef] [PubMed]
  153. Nguyen, T.; Thah, T.K.; Maclean, N.; Mahiddine, S. Mechanisms of Nucleation and Growth of Nanoparticles in Solution. Chem. Rev. 2014, 114, 7610–7630. [Google Scholar]
  154. Finney, E.; Finke, R. Nanocluster nucleation and growth kinetic and mechanistic studies, A review emphasizing transition-metal nanoclusters. J. Colloid Interface Sci. 2008, 317, 351–374. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic representation of the solid-state method and its application in environmental remediation.
Figure 1. Schematic representation of the solid-state method and its application in environmental remediation.
Ijms 23 01093 g001
Figure 2. Squematic representation of Korgel’s method. The first two steps are in solution and correspond to the preparation of the solid precursor. The third step corresponds to the solid thermolysis of the M(SC12H25)n precursor.
Figure 2. Squematic representation of Korgel’s method. The first two steps are in solution and correspond to the preparation of the solid precursor. The third step corresponds to the solid thermolysis of the M(SC12H25)n precursor.
Ijms 23 01093 g002
Figure 3. Schematic representation of the structures of Os3(CO)12 (1), Os4-(µ-H)4(CO)12 (2), and Os6(CO)18 (3). Adapted from reference [40].
Figure 3. Schematic representation of the structures of Os3(CO)12 (1), Os4-(µ-H)4(CO)12 (2), and Os6(CO)18 (3). Adapted from reference [40].
Ijms 23 01093 g003
Figure 4. TEM image of the pyrolytic product from VCl3•chitosan (1:5) and their histogram (inset) adapted from reference [72].
Figure 4. TEM image of the pyrolytic product from VCl3•chitosan (1:5) and their histogram (inset) adapted from reference [72].
Ijms 23 01093 g004
Figure 5. SEM image at two magnification levels (a,b) for Mn2O3 and their EDS analysis (c).
Figure 5. SEM image at two magnification levels (a,b) for Mn2O3 and their EDS analysis (c).
Ijms 23 01093 g005
Figure 6. XRD (a), SEM analysis (bd), EDS (e), and TEM image (fh) for the pyrolytic product from Chitosan●(FeCl3)n 1:1, hematite-Fe2O3.
Figure 6. XRD (a), SEM analysis (bd), EDS (e), and TEM image (fh) for the pyrolytic product from Chitosan●(FeCl3)n 1:1, hematite-Fe2O3.
Ijms 23 01093 g006
Figure 7. XRD pattern (a) and SEM images (b,c) for the pyrolytic product from Chitosan·(CoCl2)n.
Figure 7. XRD pattern (a) and SEM images (b,c) for the pyrolytic product from Chitosan·(CoCl2)n.
Ijms 23 01093 g007
Figure 8. SEM image (a) and EDS analysis (b) from the pyrolytic product from 1:1 Chitosan·(NiCl2)n precursor (b) and SEM image (c) and EDS analysis from (d) the 1:1 PS-co-4-PVP·(NiCl2)n precursor.
Figure 8. SEM image (a) and EDS analysis (b) from the pyrolytic product from 1:1 Chitosan·(NiCl2)n precursor (b) and SEM image (c) and EDS analysis from (d) the 1:1 PS-co-4-PVP·(NiCl2)n precursor.
Ijms 23 01093 g008
Figure 9. Luminescence spectrum of PS-4-co-PVP·AuC6F5 at several excitation wavelengths.
Figure 9. Luminescence spectrum of PS-4-co-PVP·AuC6F5 at several excitation wavelengths.
Ijms 23 01093 g009
Figure 10. SEM images for the different (AuXn)n samples: (a) AuCl, (b) AuCl3, and (c) Au(C6F5).
Figure 10. SEM images for the different (AuXn)n samples: (a) AuCl, (b) AuCl3, and (c) Au(C6F5).
Ijms 23 01093 g010
Figure 11. TEM image (a), particle size histogram, (b), TEM image in a magnified area (c), and an SEM image (d) for the pyrolytic products from the PS-co-4-PVP·(PtCl2)n precursors. Adapted from reference [99].
Figure 11. TEM image (a), particle size histogram, (b), TEM image in a magnified area (c), and an SEM image (d) for the pyrolytic products from the PS-co-4-PVP·(PtCl2)n precursors. Adapted from reference [99].
Ijms 23 01093 g011
Figure 12. SEM image for the pyrolytic precursor from PS-co-4-PVP·(ZnCl2)n in ratio 1:5. Adapted from reference [53].
Figure 12. SEM image for the pyrolytic precursor from PS-co-4-PVP·(ZnCl2)n in ratio 1:5. Adapted from reference [53].
Ijms 23 01093 g012
Figure 13. SEM image (a) and EDS analysis (b) of the SnO2 obtained from PS-co-4-PVP·(SnCl2)n precursor.
Figure 13. SEM image (a) and EDS analysis (b) of the SnO2 obtained from PS-co-4-PVP·(SnCl2)n precursor.
Ijms 23 01093 g013
Figure 14. TEM image (a) and HRTEM (b) image of pyrolytic product from the precursor PSP-co-4-PVP·(Ce(NO3)3)n.
Figure 14. TEM image (a) and HRTEM (b) image of pyrolytic product from the precursor PSP-co-4-PVP·(Ce(NO3)3)n.
Ijms 23 01093 g014
Figure 15. SEM (a,d), EDS (b), and TEM image (c) of the pyrolytic products from Eu3+ doped, Chitosan·NdCl3.
Figure 15. SEM (a,d), EDS (b), and TEM image (c) of the pyrolytic products from Eu3+ doped, Chitosan·NdCl3.
Ijms 23 01093 g015
Figure 16. HRTEM images of Au nanoparticles inside SiO2 from (PS-co-4-PVP)•(AuCl3)n•(SiO2)n precursor. Adapted from reference [94].
Figure 16. HRTEM images of Au nanoparticles inside SiO2 from (PS-co-4-PVP)•(AuCl3)n•(SiO2)n precursor. Adapted from reference [94].
Ijms 23 01093 g016
Figure 17. EDS mapping by elements of the Au/SiO2 nanocomposite from the precursor Chitosan•(AuCl3)n·(SiO2)n in 1:1 molar ratio polymer/metal. Adapted from reference [94].
Figure 17. EDS mapping by elements of the Au/SiO2 nanocomposite from the precursor Chitosan•(AuCl3)n·(SiO2)n in 1:1 molar ratio polymer/metal. Adapted from reference [94].
Ijms 23 01093 g017
Figure 18. EDS mapping for the Au/Ag//SiO2 composite from the macromolecular Chitosan·(AuCl3/AgSO3CF3)n·SiO2 precursor 1:1. Adapted from reference [94].
Figure 18. EDS mapping for the Au/Ag//SiO2 composite from the macromolecular Chitosan·(AuCl3/AgSO3CF3)n·SiO2 precursor 1:1. Adapted from reference [94].
Ijms 23 01093 g018
Figure 19. Photocatalytic activity of nanostructure metal oxides.
Figure 19. Photocatalytic activity of nanostructure metal oxides.
Ijms 23 01093 g019
Figure 20. (a) Normalized concentration changing of MB as a function of time for all TiO2 photocatalyst obtained by precursors (Chitosan)•(TiOSO4) at different temperatures. (b) Effect of pH on MB (1 × 10−5 M) discoloration using our best TiO2.
Figure 20. (a) Normalized concentration changing of MB as a function of time for all TiO2 photocatalyst obtained by precursors (Chitosan)•(TiOSO4) at different temperatures. (b) Effect of pH on MB (1 × 10−5 M) discoloration using our best TiO2.
Ijms 23 01093 g020
Figure 21. Normalized concentration changing of MB without catalyst, in the presence of a-Fe2O3·PS-co-4-PVP and in the presence of a-Fe2O3·chitosan. Adapted from reference [54].
Figure 21. Normalized concentration changing of MB without catalyst, in the presence of a-Fe2O3·PS-co-4-PVP and in the presence of a-Fe2O3·chitosan. Adapted from reference [54].
Ijms 23 01093 g021
Figure 22. Schematic diagrams for (a) energy bands of p-NiO and TiO2 before contact, (b) formation of the p-n junction and its energy diagram at equilibrium, and (c) transfer of holes from n-TiO2 to p-NiO under UV irradiation.
Figure 22. Schematic diagrams for (a) energy bands of p-NiO and TiO2 before contact, (b) formation of the p-n junction and its energy diagram at equilibrium, and (c) transfer of holes from n-TiO2 to p-NiO under UV irradiation.
Ijms 23 01093 g022
Figure 23. Schematic representation of the proposed mechanism of formation of the metal oxide nanoparticles. MXn represents the general formula of the metallic salt coordinated to the Chitosan and PSP-4-PVP polymer, }}}}}}} represents the Chitosan and PSP-4-PVP polymer. MxOy represents the respective metal oxides formed inside the graphite matrix. The given temperatures are referential general values.
Figure 23. Schematic representation of the proposed mechanism of formation of the metal oxide nanoparticles. MXn represents the general formula of the metallic salt coordinated to the Chitosan and PSP-4-PVP polymer, }}}}}}} represents the Chitosan and PSP-4-PVP polymer. MxOy represents the respective metal oxides formed inside the graphite matrix. The given temperatures are referential general values.
Ijms 23 01093 g023
Table 1. Pyrolysis temperature, phase, particle size, and morphology for the TiO2 obtained from the different precursors at several temperatures.
Table 1. Pyrolysis temperature, phase, particle size, and morphology for the TiO2 obtained from the different precursors at several temperatures.
a (Cp2TiCl2)·(Chiotosan)
Temperature (°C)PhaseSize (nm)Dispersion (nm)Morphology
500Anatase11+/−1Lamellar
600Mixture13+/−1Lamellar
700Mixture 11+/−1Porous mixture with sheets
800Rutile24+/−2Highly porous
(Cp2TiCl2)·(PS-co-4-PVP)
Temperature (°C)PhaseSize (nm)DispersionMorphology
500Anatase13+/−1Porous and sheets
600Anatase25+/−9Granular
700Anatase32+/−3Porous and granular
800Mixture 33+/−2Porous and sheet
(TiOSO4)·(Chitosan)
Temperature (°C)PhaseSize (nm)DispersionMorphology
500Anatase27+/−1Nanoparticulated microfibers
600Anatase17+/−5Nanoparticulated microfibers
700Anatase32+/−2Nanoparticulated microfibers
800Anatase7 and 32+/−1 and +/−2Nanoparticulated microfibers
(TiOSO4)·(PS-co-4-PVP)
Temperature (°C)PhaseSize (nm)DispersionMorphology
500Anatase11+/−1Irregular porous and microfibers
600Anatase20+/−1Irregular porous and microfibers
700Mixture 35+/−1Nanoparticulated microfibers
800Mixture 30+/−1Nanoparticulated microfibers
(TiO(acac)2)·(Chitosan)
Temperature (°C)PhaseSize (nm)DispersionMorphology
500Anatase12+/−1Mainly porous
600Anatase14+/−1Irregular porous and sheets
700Mixture 25+/−1Irregular porous and sheets
800Mixture 18+/−1Mainly smooth
(TiO(acac)2)·PS-co-4-PVP)
Temperature (°C)PhaseSize (nm)DispersionMorphology
500Anatase8+/−1Irregular porous
600Mixture 11+/−1Irregular porous
700Mixture 39+/−3Highly porous
800Rutile62+/−2Irregular porous and sheets
a precursor formula.
Table 2. Eg values for NiO and NiO included in the SiO2, TiO2, Al2O3, and Na4.2Ca2.8(Si6O18) matrices Adapted from reference [65].
Table 2. Eg values for NiO and NiO included in the SiO2, TiO2, Al2O3, and Na4.2Ca2.8(Si6O18) matrices Adapted from reference [65].
CompositePrecursor FormulaEg(eV)
NiOChitosan·NiCl25.2
NiOPSP-4-PVP·NiCl25.2
NiO/SiO2Chitosan·NiCl25.0
NiO/SiO2PSP-4-PVP·NiCl25.5
NiO/TiO2Chitosan·NiCl25.2
NiO/TiO2PSP-4-VP·NiCl25.2
NiO/Al2O3Chitosan·NiCl25.4
NiO/Na4.2Ca2.8(Si6O18)Chitosan·NiCl25.6
Table 3. Kinetic data for the photodegradation process of MB with NiO and NiO/SiO2, NiO/TiO2, NiO/Al2O3, and NiO/Na42Ca2.8(Si6O18) composites. Adapted from reference [65].
Table 3. Kinetic data for the photodegradation process of MB with NiO and NiO/SiO2, NiO/TiO2, NiO/Al2O3, and NiO/Na42Ca2.8(Si6O18) composites. Adapted from reference [65].
PhotocatalystDiscoloration Rate (%)
NiO-CHITOSAN71%
NiO-PS-4-PVP68%
NiO/SiO2-CHITOSAN69%
NiO/SiO2-PS-4-PVP48%
NiO/TiO2-CHITOSAN91%
NiO/TiO2-PS-4-PVP81%
NiO/Al2O3-CHITOSAN45%
NiO/Na4.2Ca2.8(Si6O18)75%
Table 4. Kinetic data for the photodegradation process of MB with ReO3 and ReO3/SiO2. Adapted from reference [62].
Table 4. Kinetic data for the photodegradation process of MB with ReO3 and ReO3/SiO2. Adapted from reference [62].
PhotocatalystPhotodegradation Rate Constant k (10−3 M·min−1)Discoloration Rate (%)
ReO3-PS-4-PVP2.864%
ReO3-Chitosan2.853%
ReO3/SiO2-PS-4-PVP3.767%
ReO3/SiO2-Chitosan1.957%
Table 5. Photocatalytic efficiency of the different ThO2 composites.
Table 5. Photocatalytic efficiency of the different ThO2 composites.
PhotocatalystDiscoloration Rate (%)
ThO2-chitosan67
ThO2-PS-4-PVP66
ThO2/SiO2-chitosan24
ThO2/SiO2-PS-4-PVP25
ThO2/TiO2-chitosan39
ThO2/TiO2-PS-4-PVP27
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Diaz, C.; Valenzuela, M.L.; Laguna-Bercero, M.Á. Solid-State Preparation of Metal and Metal Oxides Nanostructures and Their Application in Environmental Remediation. Int. J. Mol. Sci. 2022, 23, 1093. https://doi.org/10.3390/ijms23031093

AMA Style

Diaz C, Valenzuela ML, Laguna-Bercero MÁ. Solid-State Preparation of Metal and Metal Oxides Nanostructures and Their Application in Environmental Remediation. International Journal of Molecular Sciences. 2022; 23(3):1093. https://doi.org/10.3390/ijms23031093

Chicago/Turabian Style

Diaz, Carlos, Maria Luisa Valenzuela, and Miguel Á. Laguna-Bercero. 2022. "Solid-State Preparation of Metal and Metal Oxides Nanostructures and Their Application in Environmental Remediation" International Journal of Molecular Sciences 23, no. 3: 1093. https://doi.org/10.3390/ijms23031093

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

Article Metrics

Back to TopTop