Next Article in Journal
Novel Surface-Modified Bilosomes as Functional and Biocompatible Nanocarriers of Hybrid Compounds
Previous Article in Journal
Statistical Mechanics at Strong Coupling: A Bridge between Landsberg’s Energy Levels and Hill’s Nanothermodynamics
Previous Article in Special Issue
Facile Preparation of Wormlike Graphitic Carbon Nitride for Photocatalytic Degradation of Ustiloxin A
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

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

by
Carlos Diaz
1,*,
María L. Valenzuela
2,
Olga Cifuentes-Vaca
3,
Marjorie Segovia
1 and
Miguel A. Laguna-Bercero
4,*
1
Departamento de Química, Facultad de Química, Universidad de Chile, La Palmeras 3425, Nuñoa, Casilla 653, 7800003 Santiago de Chile, Chile
2
Inorganic Chemistry and Molecular Material Center, Facultad de Ingeniería, Instituto de Ciencias Químicas Aplicadas, Universidad Autónoma de Chile, Av. El Llano Subercaseaux 2801, San Miguel, 8910060 Santiago de Chile, Chile
3
Departamento Ciencias Químicas, Facultad de Ciencias Exactas, Universidad Andres Bello, Sede Concepción, Autopista Concepción-Talcahuano, 7100 Talcahuano, Chile
4
Instituto de Nanociencia y Materiales de Aragón (INMA), CSIC-Universidad de Zaragoza, 50009 Zaragoza, Spain
*
Authors to whom correspondence should be addressed.
Nanomaterials 2020, 10(12), 2470; https://doi.org/10.3390/nano10122470
Submission received: 13 November 2020 / Revised: 30 November 2020 / Accepted: 3 December 2020 / Published: 10 December 2020

Abstract

:
The medium effect of the optical and catalytic degradation of methylene blue was studied in the NiO/SiO2, NiO/TiO2, NiO/Al2O3, and NiO/Na4.2Ca2.8(Si6O18) composites, which were prepared by a solid-state method. The new composites were characterized by XRD (X-ray diffraction of powder), SEM/EDS, TEM, and HR-TEM. The size of the NiO nanoparticles obtained from the PSP-4-PVP (polyvinylpyrrolidone) precursors inside the different matrices follow the order of SiO2 > TiO2 > Al2O3. However, NiO nanoparticles obtained from the chitosan precursor does not present an effect on the particle size. It was found that the medium effect of the matrices (SiO2, TiO2, Al2O3, and Na4.2Ca2.8(Si6O18)) on the photocatalytic methylene blue degradation, can be described as a specific interaction of the NiO material acting as a semiconductor with the MxOy materials through a possible p-n junction. The highest catalytic activity was found for the TiO2 and glass composites where a favorable p-n junction was formed. The isolating character of Al2O3 and SiO2 and their non-semiconductor behavior preclude this interaction to form a p-n junction, and thus a lower catalytic activity. NiO/SiO2 and NiO/Na4.2Ca2.8(Si6O18) showed a similar photocatalytic behavior. On the other hand, the effect of the matrix on the optical properties for the NiO/SiO2, NiO/TiO2, NiO/Al2O3, and NiO/Na4.2Ca2.8(Si6O18) composites can be described by the different dielectric constants of the SiO2, TiO2, Al2O3, Na4.2Ca2.8(Si6O18) matrices. The maxima absorption of the composites (λmax) exhibit a direct relationship with the dielectric constants, while their semiconductor bandgap (Eg) present an inverse relationship with the dielectric constants. A direct relationship between λmax and Eg was found from these correlations. The effect of the polymer precursor on the particle size can explain some deviations from this relationship, as the correlation between the particle size and absorption is well known. Finally, the NiO/Na4.2Ca2.8(Si6O18) composite was reported in this work for the first time.

1. Introduction

Metal oxide nanoparticles are widely used in many applications such as coatings, catalysis, electrode materials, or sensors [1]. It is important to remark that their physical and chemical properties are strongly influenced by their agglomeration [2]. In this sense, it is well known that the incorporation of metal oxides onto inert support materials with high surface areas could help prevent particle agglomeration and also improve their reactivity and stability [3,4].
NiO is a p-type semiconductor with EG = 3.5 eV presenting multiple practical applications [4,5,6]. However, their band gap can be modified by doping with other metal oxide semiconductors, and thus changing their photocatalytic properties [5,6]. NiO has been widely used in catalysis, battery cathodes, fuel cell electrodes, electrochromic films, electrochemical supercapacitors, or magnetic materials [4,5,6]. In this sense, Bonomo et al. [7] recently reported on the electrochemical and opto-electrochemical properties of nanostructured NiO for photoconversion applications. Although these applications are determined by their band-gap, which depend on the environment [8,9], no systematic studies have been reported regarding the effect of the medium on the band-gap behavior [10,11,12]. In this sense, it is well known that the dielectric medium affects the optical properties of nanoparticles, as previously observed for Au and Ag systems [10]. The optical properties of Au nanoparticles embedded into TiO2, ZrO2, and Al2O3 have been also studied qualitatively [10]. In addition, the effect of SiO2, TiO2, and ZrO2 supports was recently analyzed showing that MoO3/SiO2 is the most efficient epoxidation catalyst [12].
The Na4.2Ca2.8(Si6O18) compound (combeite) is a crystalline phase normally obtained from the fusion of precursor Na2O⋅CaO⋅SiO2 glasses [13,14,15]. In this sense, there are no reported metal oxides using Na4.2Ca2.8(Si6O18) as a solid matrix.
In previous works, we have reported a method to prepare metal and metal oxide nanostructured materials from a thermal treatment of the Chitosan (MLn)x and PS-co-4-PVP (MLn)x macromolecular complexes [16,17,18]. The method consists of two steps: (1) Formation of both macromolecular complexes by a solvent assisted reaction between the respective polymer and the metallic salt; and (2) a thermal process of the solid under air atmosphere.
The M° and MxOy nanostructures can be easily incorporated into SiO2 matrices using a similar approach by different thermal treatments of the solid-state precursors: Chitosan (MLn)x//SiO2 and PS-co-4-PVP (MLn)x//SiO2 affording MxOY//SiO2 composites [19,20]. This method can be also used to prepare NiO//MxOy composites using SiO2, TiO2, Al2O3, or Na4.2Ca2.8(Si6O18) matrices. Although a few methods were proposed to prepare NiO/SiO2 [4,21,22], NiO/TiO2 [6,23,24,25], NiO/Al2O3 [26,27,28,29] composites, none of them is as general and simple as the one described here. As for the Na4.2Ca2.8(Si6O18) case, although this particular composition has not been reported, similar nickel oxide doped with silica matrices have been successfully synthesized via a sol–gel process [30]. Furthermore, this solid state method has been used for other systems [31]. A summary of the proposed fabrication route is shown in Figure 1 [13].
In addition, the effect of the different matrices on the optical properties will also be studied and discussed.

2. Materials and Methods

NiCl2·6H2O, tetraethyl orthosilicate (TEOS), chitosan, poly(styrene-co-4-vinilpyridine) PS-co-4-PVP, ethyl alcohol, acetic acid, and dichloromethane were supplied from Sigma-Aldrich and were used as received.

2.1. Preparation of the NiO/SiO2, NiO/TiO2, NiO/Al2O3 Composites

SiO2 was prepared according to the literature procedures [19,20]. Briefly, tetraethoxysilane (TEOS), ethanol, and acetic acid were mixed in a molar ratio of 1:4:4 with water (nanopure milli-Q), and added over the dichloromethane solution of the previously prepared chitosan (NiCl2·6H2O)x and PS-co-4-PVP (NiCl2·6H2O)x. The mixture was stirred for 3 days. The obtained gel was dried at 100 °C under a vacuum. The chitosan (NiCl2·6H2O)x//SiO2 and PS-co-4-PVP (NiCl2·6H2O)x//SiO2 precursors were finally calcined at 800 °C for 2 h under air.

2.2. Preparation of the Chitosan (NiCl2·6H2O)x//TiO2 and PS-co-4-PVP (NiCl2)x//TiO2 Precursors

TiO2 was prepared according to the literature procedures [19,20]. Briefly, titanium tetra-isopropoxide (Ti(OC3H7)4, TTIP) ethanol and acetic acid were mixed in a molar ratio of 1:4:4 with water (nanopure milli-Q), and added over the dichloromethane solution of the previously prepared chitosan (NiCl2·6H2O)x and PS-co-4-PVP (NiCl2·6H2O)x. The mixture was stirred for 3 days. The obtained gel was dried at 100 °C under a vacuum. The solid chitosan (NiCl2·6H2O)x//TiO2 and PS-co-4-PVP (NiCl2·6H2O)x//TiO2 precursors were calcined at 800 °C for 2 h under air.

2.3. Preparation of the Chitosan (NiCl2·6H2O)x//Al2O3 and PS-co-4-PVP (NiCl2)x//Al2O3 Precursors

Al2O3 was prepared according to the literature procedures [27,28,29,30]. Briefly, AlCl3, ethanol, and acetic acid were mixed in a molar ratio of 1:4:4 with water (nanopure milli-Q), and added over the dichloromethane solution of the previously prepared chitosan (NiCl2·6H2O)x and PS-co-4-PVP (NiCl2·6H2O)x. The mixture was stirred for 3 days. The obtained gel was dried at 100 °C under a vacuum. The solid chitosan (NiCl2·6H2O)x//Al2O3 and PS-co-4-PVP (NiCl2·6H2O)x//Al2O3 precursors were calcined at 800 °C for 2 h under air.

2.4. Preparation of the Precursors: Chitosan (NiCl2·6H2O)x//NiO/Na4.2Ca2.8(Si6O18) and PS-co-4-PVP (NiCl2)x//NiO/Na4.2Ca2.8(Si6O18)

The compounds were prepared according to the literature procedures [28]. Briefly, tetraethoxysilane (TEOS), ethanol, and acetic acid were mixed in a molar ratio of 1:4:4 with water (nanopure milli-Q), then Na2O, CaO, and SiO2 solids (in mol% of 14:1.5:73) were added over the dichloromethane solution of the previously prepared chitosan (NiCl2·6H2O)x and PS-co-4-PVP (NiCl2·6H2O)x. The mixture was stirred for 3 days. The obtained gel was dried at 100 °C under a vacuum. The solid chitosan (NiCl2·6H2O)x//Na2O CaO SiO2 and PS-co-4-PVP (NiCl2·6H2O)x//Na2O CaO SiO2 precursors were calcined at 800 °C for 2 h under air.
The coordination of the polymer was confirmed by IR analysis, as the broad ν(OH)+ ν(NH) band observed at 3448 cm−1 for free chitosan becomes unfolded upon coordination, shifting in the range of 3345–3393 cm−1. On the other hand, the ν(py) band is shifting to high frequencies upon coordination [16,17,18].
Finally, polymer-metal complexes were placed into a box furnace (lab tech) using a pyrolysis temperature of 180 °C for the precursor complexes and 800 °C for the polymer complexes. Additional experimental conditions are summarized in Table 1.

2.5. Characterization

IR spectra were recorded with a FT-IR Jasco 4600 spectrophotometer (Jasco Inc., Easton, MD, USA). Scanning electron microscopy (SEM) was performed on a JEOL 5410 scanning electron microscope (JEOL Ltd., Tokyo, Japan). Elemental microanalysis was performed by energy dispersive X-ray (EDS) analysis using a NORAN Instrument micro-probe attached to the SEM (Thermo Scientific, Waltham, MA, USA). High-resolution transmission electron microscopy (HR-TEM) was performed using a JEOL 2000FX TEM microscope (JEOL Ltd., Tokyo, Japan)at 200 kV to characterize the average particle size, distribution, and elemental and crystal composition. EDS analysis was performed in individual particles in order to discriminate NiO from the matrix. Average particle sizes were calculated using the Digital Micrograph software (Gatan, Inc., Pleasanton, CA, US). Methylene blue (MB) was used as a model compound to test the photocatalytic properties at 655 nm under UV-Vis illumination (Shimadzu UV-2600 spectrophotometer, Shimadzu Coorporation, Kyoto, Japan) using a xenon lamp (150 W) positioned 20 cm away from the photoreactor in a 330–680 nm range at room temperature, to avoid the self-degradation and thermal catalytic effects of cationic dye. Suspensions were stirred in the dark for 60 min to establish an adsorption/desorption equilibrium, after which the photocatalytic discoloration of MB was initiated.

3. Results and Discussion

3.1. Composite NiO/SiO2

The X-ray diffraction pattern of the as-synthesized NiO/SiO2 composite for the material from the chitosan precursor is shown in Figure 2a. All the reflection peaks of the XRD pattern can be indexed to NiO and SiO2 phases [19] (JPDS no. 03-065-2901 for NiO and JPDS no. 01-088-1535 for SiO2). The broad feature appearing at 22° corresponds to amorphous silica [19]. Similar X-ray diffraction patterns for NiO from the PVP precursor were obtained.
The SEM analysis (Figure 2b) shows irregular particle agglomerates, as typically observed from the preparation of nanoparticles using the solid-state thermal route [30]. From the TEM analysis, the agglomeration of NiO nanoparticles embedded into a mesh of SiO2 can be observed in Figure 2c, where these agglomerates are composed of fused NiO nanoparticles. The size of these nanoparticles are in the range of 14 nm with a mean size of 25 nm (Figure 2c). Detailed HR-TEM images in Figure 2e,f show a homogeneous dispersion of NiO over the silica network. However, it was not possible to acquire high resolution images in order to study the interfaces between NiO and the different matrices. In any case, as also confirmed by SEM-EDS mapping (Figure 2g), there is a uniform distribution of NiO and SiO2 particles. Similar results were observed for NiO obtained from the PVP precursor (see Supplementary Materials, Figure S1). The only difference is that NiO particles are bigger in size ca. 100 nm.

3.2. NiO/TiO2

Figure 3 shows the XRD pattern of the NiO/TiO2 nanocomposite from the chitosan precursor, where the anatase phase and NiO are observed as single phases. Using this method, the pure TiO2 anatase phase was obtained, in contrast with other solution methods, where a mixture of anatase and rutile in the NiO/TiO2 composite was obtained [22]. The NiO/TiO2 composite shows a “cotton” type morphology from the chitosan precursor (Figure 3b), whereas the morphology from the PVP precursor presents a more densified structure, as shown in Figure 3c. The SEM-EDS mapping, shown in Figure 2g, indicates an homogeneous distribution of NiO and TiO2. Similar results were obtained for the NiO/TiO2 from the PVP precursor (see Supplementary Materials, Figure S2).
The TEM analysis (Figure 3d,e) presents a “spider web” TiO2 network where the NiO nucleates forming agglomerated nanoparticles. They present a mean particle size of 25 nm (Figure 2f). A similar TEM analysis was observed for NiO/TiO2 obtained from the PVP precursor (Figure 3b and Supplementary Materials, Figure S2).

3.3. NiO/Al2O3

Figure 4a shows the XRD pattern of the NiO/Al2O3 composite from the chitosan precursor where the corresponding peaks of γ-Al2O3 and NiO can be observed.
The effect of the polymer template on the morphology can be observed in Figure 4b,c. The chitosan precursor induces a “cotton” type morphology, while the PVP precursor also combines dense and irregular zones. Figure 4f shows an elemental mapping image demonstrating that NiO is well dispersed inside Al2O3. A complete characterization is shown in Supplementary Materials, Figure S3.
As observed for the NiO/TiO2 system, the TEM analysis (Figure 4e) shows a “spider web” network of Al2O3 where the NiO nucleates form agglomerates. The histogram (Supplementary Materials, Figure S3) shows a particle mean size of 17 nm. The HRTEM image of the NiO/Al2O3 from the PVP precursor is shown in Supplementary Materials, Figure 3c, where it can be observed that the medium particle size is 32 nm.

3.4. NiO/Na4.2Ca2.8(Si6O18)

The XRD pattern of the NiO/Na4.2Ca2.8(Si6O18) composite prepared from the chitosan precursor indicates the formation of NiO inside the glass Na4.2Ca2.8(Si6O18) (see Figure 5a). The XRD pattern is in agreement with those reported in the literature [13,14,15]. The observed morphology is similar to the one previously reported [13,14,15] (see Figure 5b,c), also presenting a uniform distribution of NiO inside the Na4.2Ca2.8(Si6O18) (Figure 5d). Similar conclusions can be deduced for the PVP precursor (see Supplementary Materials, Figure S4).
A summary of the medium particle sizes for NiO included into the different matrices is presented in Table 2, where the effect of the matrix and that of the polymer precursors on the final particle sizes can be observed.
The nanoparticle size of NiO obtained from the PVP precursor inside the matrices follow the order of SiO2 > TiO2 > Al2O3, while that for the NiO from the chitosan precursor does not present a significant effect on the nanoparticle size.

3.5. Photocatalytic Behavior

Although the main applied property of NiO is in the field of electrochemistry as Li-ion batteries [32] and supercapacitors applications, [33] its application as a photocatalytic activity toward organic dyes have also been suggested [34]. In any case, reports on the photocatalytic activity toward organic dyes using NiO/matrices are scarce. Yu et al. [6] found a higher photocatalytic activity for NiO/TiO2 than for pure NiO, towards the photodegradation of p-chlorophenol. Regarding the photocatalytic efficiency when using composites, important parameters to be considered include the formation of hierarchical porous structures, the dispersion of the catalytic semiconductor on the matrix surface, and the p-n junction in a NiO/MxOy composite, where a new band gap will be formed with a most favorable value for the photodegradation chemical processes.

3.6. NiO

Methylene blue (MB) is extensively used as an organic dye in coloring paper, temporary hair colorant, dyeing cottons, and coating for paper stock [35]. The removal of this hazardous dye is considered as one of the growing requirements in recent years. The photocatalytic experiments were carried on the sample with definite dye concentration under dark conditions and UV irradiation. The band-gap of the NiO is 5.0 and 5.2 eV, when it is prepared from chitosan and PVP precursors, respectively. For the semiconductor metal oxides, their band gap value dictates their photocatalytic activity [35,36]. For this reason, the band gap of the C3–C8 composites was determined. These values are: 5.0, 5.2, and 5.4 eV for the NiO/SiO2, NiO/TiO2, NiO/Al2O3 composites, respectively, all obtained from the chitosan precursors. The values for the PVP precursor are: 5.5 eV, 5.2 eV for the NiO/SiO2, NiO/TiO2 composites, respectively. Those values do not change significantly, and are slightly higher than those reported previously, which can be due to their bigger particle sizes [34] (see Supplementary Materials, Figure S5).
The changes in the absorption spectra of the MB aqueous solution exposed to UV light for various times in the presence of NiO are shown in Supplementary Materials, Figure S6. The peak at 655 nm is characteristic of methylene blue and decreases with the irradiation time. Figure 6 shows the plot of time vs. concentration of methylene blue measured as C/Co for NiO arising from both precursors, obtaining a catalytic efficiency of ~68% and ~71% of degradation in 5 h (see Figure 6c). Both degradation processes follow a zero order, as shown in Figure 6b,d. As previously mentioned, only a few photodegradation studies for NiO have been reported. For example, using 3 nm NiO nanoparticles [34] and NiO nanofibers [5], a moderated catalytic activity towards Rhodamine B was observed. In both cases, the degradation kinetic was zero order, which means that the rate of degradation does not depend on the MB concentration. This type of model is normally observed when the surface of the photocatalyst is saturated with the dye, so that the degradation rate remains relatively constant, depending only on the generation of photo-induced charges in the catalyst.

3.7. NiO in Matrices

The photocatalytic activity towards MB degradation for the NiO composite using different matrices is shown in Figure 7. The degradation rate of the NiO/TiO2 composite is shown for comparison. In any case, the photocatalytic activity of these NiO compounds is still far from the pure TiO2 standard phase [37]. For example, we have recently reported a 98% discoloration rate in only 25 min for TiO2 nanostructures using similar synthetic routes, and the degradation of commercial TiO2 (Degussa P25) is about 75% of MB under the same experimental conditions [38]. A representative plot of MB absorption at 655 nm vs. time is given in Supplementary Materials, Figure S6. A summary of the kinetic degradation data is also displayed in Table 3.
As seen in Figure 6, the NiO from the chitosan precursor produces a higher activity than that arising from the PVP precursor. These results also apply for both SiO2 and TiO2 matrices. Interestingly, the most efficient photocatalytic activity was observed for the NiO/TiO2 composite with a 91% degradation of methylene blue in 5 h. This can be probably related with a matrix effect of SiO2, TiO2, and Al2O3.
Our results of catalytic degradation for the NiO//TiO2 composite (about 91%) are similar or slightly higher than those reported in the literature. Ahmed claimed 90% of catalytic degradation efficiency on the NiO//TiO2 composite prepared from titanium chloride and nickel acetylacetonate [39]. Faisal et al. obtained a similar catalytic degradation efficiency using an ultrasonication method [40]. Sim et al. reported 86% of the degradation efficiency using plasma enhanced chemical vapor deposition (PECVD) with hydro-oxygenated amorphous titanium dioxide obtained from titanium tetra-isopropoxide [Ti(OC3H7)4, TTIP] liquid as a precursor [41]. Finally, Chen et al. reported 86% catalytic degradation of MB using a method that involves incipient wet impregnation of the nickel oxide (NiO) nanoparticles over previously prepared TiO2 nanotubes [24].
It is suggested that for the most catalytically active TiO2 as the matrix, a p-n junction can be formed acting NiO as p-NiO and TiO2 as n-TiO2, see Supplementary Materials, Figure S7, leading to a reduction of the recombination rate of photogenerated electron-hole pairs, which is known to enhance the photocatalytic activity of TiO2. A detailed description of the mechanism can be found on Supplementary Materials, Figure S8. 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. On the other hand, the less efficient photocatalyst toward MB degradation arises probably from an insulating Al2O3 effect [28,42], which preclude the p-NiO behavior. This is in agreement with the observed photocatalytic decrease for the TiO2/SiO2 composite in comparison with pure TiO2. In the case of the NiO/SiO2 composite, the lower photocatalytic activity is probably a consequence of the high porous morphology which is induced by the SiO2 matrix. All the photodegradation processes of MB with NiO/SiO2, NiO/TiO2, NiO/Al2O3, and NiO/Na4.2Ca2.8(Si6O18) composites exhibited a zero order kinetic law, as shown in Supplementary Materials, Figure S9.

3.8. Photocatalytic Activity of the NiO/Na4.2Ca2.8(Si6O18) Composite

The photocatalytic activity of the NiO/Na4.2Ca2.8(Si6O18) composite obtained from the chitosan precursor is shown in Figure 7 and the kinetic data is also shown in Table 2. It is observed that the photocatalytic activity is higher than that of NiO, NiO/SiO2, and NiO/Al2O3 but lower than of NiO/TiO2. It is concluded that the Na4.2Ca2.8(Si6O18) sample presents a similar behavior to those of the SiO2 sample.

3.9. Effect of the Matrices on λmax and Eg

Figure 8 shows the variation of both Eg and λmax for the different matrices. The respective UV-Vis absorption spectra of the composites are shown in Supplementary Materials, Figure S9. The band-gap values were estimated from these spectra using the Tauc procedure (Supplementary Materials, Figure S9). Considering that the static dielectric constants (K) for the matrices are: SiO2 3.9; TiO2 80, and Al2O3 8.8 [43], both Eg and λmax could be related to the dielectric constant of the matrix. Unfortunately, there is no available data for Na4.2Ca2.8(Si6O18). The dependence of Eg with the dielectric constant ε is not totally understood, where several relationships have been previously found [43,44,45]. The shape of the experimental or theoretical expression depends, among others, on the type of materials. On the other hand, the relationship of λmax with ε and the refractive index n is known for metallic nanoparticles [8]:
λ max   α   λ p 2 ε + 1 2   λ p   n
However, the analogue relationship for metal oxides is not completely understood. The plot of λmax for the NiO vs. the refractive index [46,47] for the SiO2, TiO2, and Al2O3 matrices shows an inverse and irregular relationship (see Supplementary Materials, Figure S10). Then, according to Figure 8, the variations of Eg and λmax can be explained by a physical effect of the medium reflected in their dielectric constant of the different matrices. A close inspection of Figure 8 suggests the presence of three linear trends. In Figure 8, it can be observed that λmax varied inversely with the properties of the matrices (i.e., dielectric constant for instance) for the NiO obtained from both polymers in an approximate linear behavior. The composite C8 having a “silica like” matrix does not follow this trend due to an unknown effect. Although the dependence of λmax with the dielectric or refractive index given by Equation (1) indicates a direct linear dependence, our results show an inverse linear trend. Then, the Equation (1) may not be valid for metallic oxides. A new equation is proposed (Equation (2), curve a in Figure 8), which could arise from the general trends for nanostructured metallic oxides. This is consistent with the fact observed in Supplementary Materials, Figure S9, where an inverse relationship of λmax with n is shown.
In addition, Eg values vary in a direct or inverse way depending on the NiO polymer precursor (direct behavior for the chitosan; curve b, Equation (3) or inverse for the PVP precursor; curve c, Equation (4)). As previously mentioned, the dependence of Eg with the dielectric constant ε is not totally understood, and this is a matter of controversy in the literature. The inverse relationship (curve c) is in agreement with the results reported by Hervé and Vandamme [48], while the direct relationship (curve b) shows a similar trend to that shown by Kumar and Singh [49]. In any case, we do not have any clear explanation of the different dependencies of Eg with n and ε when using the different polymer precursors.
From the plot shown in Figure 8, the following equations can be established:
λmax = a/(ε,n); valid for NiO from chitosan and PVP
Eg = b/(ε,n); valid for NiO from chitosan
Eg = c/(ε,n); valid for NiO from PVP
The following equations are then obtained by combining both expressions:
Eg = ab/λmax; valid for NiO from chitosan
Eg = cλmax/a; valid for NiO from PVP
In agreement with these new expressions, we can explain the effect of the physical properties of the matrices on the band gap with the refraction index or the dielectric constant. The experimental data fits into these equations, as seen in Figure 9 plots d (Equation (5)) and e (Equation (6)). These equations describing the effect of the medium modulated by various solid matrices on the band gap and the maximum absorption could be valid for other nanostructured metal oxides included in solid matrices. In order to validate this, additional experiments with other systems are being carried out.
Experiments linking the band gap with the size and the maxima absorption of nanoparticles have been performed for other metal oxides such as ZnO [50,51], as well as for noble metal nanoparticles such as Au [52], Ag [32], and Pt [53]. However, there are no studies in the literature about the medium expressed by solid matrices on nanostructured metallic oxides.

4. Conclusions

NiO/SiO2, NiO/TiO2, NiO/Al2O3, and NiO/glass composites were satisfactorily prepared by a solid-state synthesis from the chitosan and PVP precursors. XRD, SEM/EDS, and HR-TEM were used to characterize the new formed composites. It was concluded that the nature of the precursor polymer influences the morphology, as well as the size of the obtained nanoparticles. The chitosan precursor induces the smallest NiO nanoparticles and also their respective nanocomposites. In addition, the nature of the matrix influences the NiO nanoparticle size, following the order of SiO2 > TiO2 > Al2O3 for the PVP precursor. However, no relationship on the particle size was observed for the NiO obtained from the chitosan precursor.
The efficiency on the photocatalytic activity depends on the formation of a p-n junction between NiO acting as p-NiO and the metal oxide matrix acting as n-metal oxide. TiO2 presents the most effective p-NiO//n-TiO2 junction. On the other hand, the optical parameters Eg and λmax depends on the dielectric constant and the refractive index of the matrix medium in a manner which depends on the preparation procedure. The “silica like” Na4.2Ca2.8(Si6O18) matrix does not follow these correlations. New equations describing the effect of the physical properties (dielectric constant and the refractive index) are proposed, which could be used for other metal oxides included in solid matrices.

Supplementary Materials

The following are available online at https://www.mdpi.com/2079-4991/10/12/2470/s1, Figure S1: (a) SEM-EDS mapping by element and (b) TEM images and (c) HRTEM image of NiO from the PS-co-4-PVP (NiCl2)x //SiO2 precursor, Figure S2: (a) SEM-EDS mapping by element from the PS-co-4-PVP⋅(NiCl2)n)x//TiO2 precursor and (b) TEM images and its histogram of the pyrolytic product from the PS-co-4-PVP⋅( NiCl2)n)x//TiO2 precursor, Figure S3: (a) SEM-EDS mapping by element and (b) TEM images, Histogram and electron diffractions, of the pyrolytic product from the precursor PS-co-4-PVP⋅(NiCl2)n)x //Al2O3 and (c) Histogram of TEM image of NiO from Chitosan⋅(NiCl2)n)x //Al2O3 precursor, Figure S4: SEM image and EDS mapping analysis for the composite NiO/Na4.2 Ca2.8 (Si6O18) from the PS-co-4-PVP⋅(NiCl2)n)//SiO2⋅CaO⋅Na2O precursor, Figure S5: Tauc determination of Eg values for NiO from Chitosan⋅(NiCl2)n. and PS-co-4-PVP⋅(NiCl2)n and for NiO/SiO2, NiO/TiO2, NiO/Al2O3 and NiO/Na4.2 Ca2.8(Si6O18) composites, Figure S6: Absorbance vs time for the blue methylene degradation for NiO (a) and for the composites NiO/SiO2, NiO/TiO2, NiO/Al2O3; (b) NiO/SiO2 from Chitosan⋅(NiCl2)x//TiO2; (c) NiO/SiO2 from PS-co-4-PVP⋅(NiCl2)x//TiO2; (d) NiO/TiO2 from Chitosan⋅(NiCl2)x //TiO2; (e) NiO/TiO2 from PS-co-4-PVP⋅(NiCl2)x //TiO2; (f) NiO/Al2O3 from Chitosan⋅(NiCl2)x //Al2O3; (g) NiO/ Na4.2 Ca2.8 (Si6O18) from Chitosan⋅(NiCl2)x // Na2O⋅CaO⋅SiO2, Figure S7: Schematic diagrams for (a) energy bands of p-NiO and TiO2 before contact, (b) formation of p-n junction and its energy diagram at equilibrium and (c) transfer of holes from n-TiO2 to p-NiO under UV irradiation, Figure S8: The photodegradation mechanism of NiO/TiO2 composites, Figure S9: Kinetic plot of the blue methylene degradation with the composites NiO/SiO2, NiO/TiO2, NiO/Al2O3 and NiO/Na4.2Ca2.8(Si6O18), Figure S10: UV-Vis absorption spectra of the composites, Figure S11: Plot of λmax for the NiO vs the refractive index for the matrices SiO2, TiO2 and Al2O3.

Author Contributions

C.D. and M.L.V. conceived the idea and designed the experiments. O.C.-V. and M.S. carried out the preparation of the nanoparticles and performed the experimental analysis. M.A.L.-B. performed the TEM experimental measurements and analysis. C.D., M.L.V., M.S., O.C.-V. and M.A.L.-B. analyzed the data and wrote the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

The authors acknowledge the Fondecyt Project 1160241 for financial support. MLB also acknowledges the project PID2019-107106RB-C32 funded by the Spanish Ministry of Science and Innovation. The use of Servicio General de Apoyo a la Investigación (SAI, University of Zaragoza) is finally acknowledged.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fernandez-Garcia, E.; Martinez-Arias, A.; Hanson, J.C.; Rodriguez, J.A. Nanostructured oxides in chemistry: Characterization and properties. Chem. Rev. 2014, 104, 4063–4104. [Google Scholar] [CrossRef] [PubMed]
  2. Srivastava, A.K. Oxide Nanostructures, Growth, Macrostructures and Properties; Pan Stanford Publishing: Singapore, 2014. [Google Scholar]
  3. Liu, S.; Han, M.-Y. Silica-coated metal nanoparticles. Chem. Asian J. 2010, 5, 36–45. [Google Scholar] [CrossRef]
  4. Sietsma, J.R.A.; Meeldijk, J.D.; Den Breejen, J.P.; Versluijs-Helder, M.; van Dillen, J.A.; De Jongh, P.E.; De Jong, K.P. The Preparation of Supported NiO and Co3O4 Nanoparticles by the Nitric Oxide Controlled Thermal Decomposition of Nitrates. Angew. Chem. Int. Ed. 2007, 46, 4547–4549. [Google Scholar] [CrossRef] [PubMed]
  5. Zhang, Z.; Shao, C.; Li, X.; Wang, C.; Zhang, M.; Liu, Y. Electrospun Nanofibers of p~Type NiO/n~Type ZnO Heterojunctions with Enhanced Photocatalytyc Activity. ACS Appl. Mater. Interface Sci. 2010, 10, 2915–2923. [Google Scholar] [CrossRef] [PubMed]
  6. Yu, J.; Wang, W.; Cheng, B. Synthesis and Enhanced Photocatalytic Activity of a Hierarchical Porous Flowerlike p–n Junction NiO/TiO2 Photocatalyst. Chem. Asian 2010, 5, 2499–2506. [Google Scholar] [CrossRef] [PubMed]
  7. Bonomo, M.; Dini, D.; Decker, F. Electrochemical and Photoelectrochemical Properties of Nickel Oxide (NiO) with Nanostructured Morphology for Photoconversion Applications. Front. Chem. 2018, 6, 601. [Google Scholar] [CrossRef] [PubMed]
  8. Jensen, T.R.; Duval, M.L.; Kelly, K.L.; Lazarides, A.A.; Schatz, G.C.; Van Duye, R.P. Nanosphere Lithography:  Effect of the External Dielectric Medium on the Surface Plasmon Resonance Spectrum of a Periodic Array of Silver Nanoparticles. J. Phys. Chem. 1999, 103, 9846–9853. [Google Scholar] [CrossRef]
  9. Kelly, K.L.; Coronado, E.; Zhao, L.L.; Schatz, G. The optical properties of metal nanoparticles: The influence of size, shape, and dielectric environment. J. Phys. Chem. 2003, 107, 668–677. [Google Scholar] [CrossRef]
  10. Prevel, B.; Palpant, B.; Lerme, J.; Pellarin, M.; Treilleux, M.; Saviot, L.; Duval, E.; Perez, A.; Broyer, M. Comparative Analysis of Optica Properties of Gold and Silver Clusters Embedded in Alumina Matrix. NanoStructured Mater. 1999, 12, 307–310. [Google Scholar] [CrossRef]
  11. Matsuoka, J.; Yoshida, H.; Nasu, H.; Kamiya, K. Preparation of Gold Microcrystal-Doped TiO2, ZrO2 and Al2O3 Films through Sol-Gel Process. J. Sol Gel. Tech. 1997, 9, 145–155. [Google Scholar] [CrossRef]
  12. Chandra, P.; Doke, D.S.; Unbarkar, S.D.; Biradar, A.V. One-pot synthesis of ultrasmall MoO3 nanoparticles supported on SiO2, TiO2, and ZrO2 nanospheres: An efficient epoxidation catalyst. J. Mater. Chem. A 2014, 2, 19056–19060. [Google Scholar] [CrossRef]
  13. Soares, V.O.; Daguano, J.K.M.B.; Lombello, C.B.; Bianchin, O.S.; Gonçalves, L.M.G.; Zanotto Edgar, D. New sintered wollastonite glass-ceramic for biomedical applications. Ceram. Int. 2018, 44, 20019–20027. [Google Scholar] [CrossRef]
  14. Volzone, C.; Stábile, F.M. Structural Changes by Thermal Treatment up to Glass Obtention of P2O5-Na2O-CaO-SiO2 Compounds with Bioglass Composition Types. New J. Glass Ceram. 2013, 3, 53–57. [Google Scholar] [CrossRef] [Green Version]
  15. Martel Estrada, S.A.; Armendáriz, I.O.; Torres García, A.; Hernández, J.F.; Rodríguez, C.A. González Evaluation of In Vitro Bioactivity of 45S5 Bioactive Glass/Poly Lactic Acid Scaffolds Produced by 3D Printing. Int. J. Compos. Mater. 2017, 7, 144–149. [Google Scholar]
  16. Diaz, C.; Barrientos, L.; Carrillo, D.; Valdebenito, J.; Valenzuela, M.L.; Allende, P.; Geaney, H.; O’Dwyer, C. Solvent-less method for efficient photocatalytic α-Fe2O3 nanoparticles using macromolecular polymeric precursors. New J. Chem. 2016, 40, 6768–6776. [Google Scholar] [CrossRef]
  17. Diaz, C.; Valenzuela, M.L.; Laguna-Bercero, 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 metal macromolecular complex precursors. RSC Adv. 2017, 7, 27729–27736. [Google Scholar] [CrossRef] [Green Version]
  18. Díaz, C.; Valenzuela, M.L.; Laguna, A.; Lavayen, V.; Jimenez, J.; Power, L.; O’Dwyer, C. Metallophosphazene Precursor Routes to Solid-State Deposition of Metallic and Dielectric Micro- and Nanostructures on Si and SiO2. Langmuir 2010, 26, 10223–10233. [Google Scholar] [CrossRef]
  19. Díaz, 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]
  20. Díaz, C.; Valenzuela, M.L.; Bobadilla, D.; Laguna-Bercero, M.A. Bimetallic Au//Ag Alloys inside SiO2 using a solid-state method. J. Clust. Sci. 2017, 28, 2809–2815. [Google Scholar] [CrossRef] [Green Version]
  21. Zong, J.; Zhu, Y.; Yang, X.; Li, C. Confined growth of CuO, NiO, and Co3O4 nanocrystals in mesoporous silica (MS) spheres. J. Alloys Compd. 2011, 509, 2970–2975. [Google Scholar] [CrossRef]
  22. Shufu, C.; Sujuan, Z.; Wei, L.; Wei, Z. Preparation and activity evaluation of p-n junction photocatalyst NiO/TiO2. J. Hazard. Mater. 2008, 155, 320–326. [Google Scholar] [CrossRef] [PubMed]
  23. Yu, J.H.; Yang, H.; Jung, R.H.; Lee, J.W.; Boo, J.H. Hierarchical NiO/TiO2 composite structures for enhanced electrochromic durability. Thin Solid Films 2018, 664, 1–5. [Google Scholar] [CrossRef]
  24. Chen, J.-Z.; Chen, T.-H.; Lai, L.-W.; Li, P.-Y.; Liu, H.-W.; Hong, Y.-Y.; Liu, D.-S. Preparation and Characterization of Suface Photocatalytic Activity with NiO/TiO2 Nanocomposite Structure. Materials 2015, 8, 4273–4286. [Google Scholar] [CrossRef] [Green Version]
  25. Wang, T.T.; Chiang, C.L.; Lin, Y.C.; Srinivasadesikan, V.; Lin, M.C.; Lin, Y.G. KSCN-activation of hydrogenated NiO/TiO2 for enhanced photocatalytic hydrogen evolution. Appl. Surf. Sci. 2020, 511, 145548. [Google Scholar] [CrossRef]
  26. Makhlouf, S.; AKhalil, K.M.S. Humidity sensing properties of NiO/Al2O3 nanocomposite materials. Solid State Ion. 2003, 164, 97–106. [Google Scholar] [CrossRef]
  27. Rogojan, R.; Andronescu, E.; Ghitulica, C.; Vasile, B. Synthesis and Characterization of Alumina Nano-power Obtanined by Sol-Gel Method. UPB Sci. Bull. Ser. B 2011, 73, 67–76. [Google Scholar]
  28. Adams, L.A.; Essien, E.R.; Shaibu, R.O.; Oki, A. Sol-Gel Synthesis of SiO2-CaO-Na2O-P2O5 Bioactive Glass Ceramic from Sodium Metasilicate. New J. Glass Ceram. 2013, 3, 11–15. [Google Scholar] [CrossRef] [Green Version]
  29. Kamire, R.J.; Majewski, M.B.; Hoffeditz, W.L.; Ohelan, B.T.; Farha, O.K.; Hupop, J.T.; Wasielewski, M.R. Photodriven hydrogen evolution by molecular catalysts using Al2O3-protected perylene-3,4-dicarboximide on NiO electrodes. Chem. Sci. 2017, 8, 541–549. [Google Scholar] [CrossRef] [Green Version]
  30. El-Shaarawy, M.; El-rafa, M.A.; Gouda, M.; Khoder, H.; Ramadan, M. Electrical and Optical Properties for Nano (SiO2)100-x:(NiO)x Glass Matrix. IOSR J. Appl. Phys. 2014, 6, 18–29. [Google Scholar] [CrossRef]
  31. 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]
  32. Wang, X.; Li, X.; Sun, X.; Li, F.; Wang, Q.; He, D. Nanostructured NiO electrode for high rate Li-ion batteries. J. Mater. Chem. 2011, 21, 3571–3573. [Google Scholar] [CrossRef]
  33. Dar, F.I.; Moonooswamy, K.R.; Es-Souni, M. Morphology and property control of NiO nanostructures for supercapacitor applications. Nanoscale Res. 2013, 8, 1–7. [Google Scholar] [CrossRef] [Green Version]
  34. Duan, H.; Zheng, X.; Yuan, S.; Li, Y.; Tian, Z.; Deng, Z.; Su, B. Sub-3 nm NiO nanoparticles: Controlled synthesis, and photocatalytic activity. Mater. Lett. 2012, 81, 245–247. [Google Scholar]
  35. 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]
  36. 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]
  37. Carlucci, C.; Xu, H.; Scremin, B.F.; Giannini, C.; Sibillano, T.; Carlino, E.; Videtta, V.; Gigli, G.; Ciccarella, G. Controllable one-pot synthesis of anatase TiO2 nanorods with the microwave-solvothermal method. Sci. Adv. Mater. 2014, 6, 1668–1675. [Google Scholar] [CrossRef]
  38. Allende-González, P.; Laguna-Bercero, M.A.; Barrientos, L.; Valenzuela, M.L.; Díaz, C. Solid State Tuning of TiO2 Morphology, Crystal Phase, and Size through Metal Macromolecular Complexes and Its Significance in the Photocatalytic Response. ACS Appl. Energy Mater. 2018, 7, 3159–3170. [Google Scholar] [CrossRef]
  39. Ahmed, M.A. Synthesis and structural features of mesoporous NiO/TiO2 nanocomposites prepared by sol–gel method for photodegradation of methylene blue dye. J. Photochem. Photobiol. A Chem. 2012, 238, 63–70. [Google Scholar] [CrossRef]
  40. Faisal, M.; Harraza, F.A.; Ismailc, A.A.; El-Tonib, A.M.; Al-Sayaria, S.A. Novel mesoporous NiO/TiO2 nanocomposites with enhanced photocatalytic activity under visible light illumination. Ceram. Int. 2018, 44, 7047–7056. [Google Scholar] [CrossRef]
  41. Sim, L.C.; Ng, K.W.; Ibrahim, S.; Saravanan, P. Preparation of Improved p-n Junction NiO/TiO2 Nanotubes for Solar-Energy-Driven Light Photocatalysis. Int. J. Photoenergy 2013, 2013. [Google Scholar] [CrossRef] [Green Version]
  42. Gangwar, J.; Gupta, B.K.; Tripathi, S.K.; Srivastava, A.K. Phase dependent thermal and spectroscopic responses of Al2O3 nanostructures with different morphogenesis. Nanoscale 2015, 7, 13313–13344. [Google Scholar] [CrossRef]
  43. Ravichandran, R.; Wang, A.; Wager, J. Solid state dielectric screening versus band gap trends and implications. Opt. Mater. 2016, 60, 181–187. [Google Scholar] [CrossRef] [Green Version]
  44. Chadi, D.J.; Cohen, M.L. Correlation between the static dielectric constant and the minimum energy gap. Phys. Lett. A 1974, 49, 381–382. [Google Scholar] [CrossRef] [Green Version]
  45. Kalyanaraman, S.; Shajinshinu, P.M.; Vijayalakshmi, S. Refractive index, band gap energy, dielectric constant and polarizability calculations of ferroelectric Ethylenediaminium Tetrachlorozincate crystal. J. Phys. Chem. Solid 2015, 86, 108–113. [Google Scholar] [CrossRef]
  46. Ghazal, M.N.; Deparis, O.; De Coninck, J.; Gaigneaux, E.M. Tailored refractive index of inorganic mesoporous mixed-oxide Bragg stacks with bio-inspired hygrochromic optical properties. J. Mater. Chem. C 2013, 1, 6202–6209. [Google Scholar] [CrossRef]
  47. Kischkat, J.; Peters, S.; Gruska, B.; Semtsiv, M.; Chashnikova, M.; Klinkmuller, M.; Fedosenko, O.; Machulik, S.; Alesksandrova, A.; Monastyrski, G.; et al. Mid-infrared optical properties of thin films of aluminum oxide, titanium dioxide, silicon dioxide, aluminum nitride, and silicon nitride. Appl. Opt. 2012, 51, 6789–6798. [Google Scholar] [CrossRef]
  48. Hervé, P.; Vandamme, L.K.J. General relation between refractive index and energy gap in semiconductors. Infrared Phys. Technol. 1994, 35, 609–615. [Google Scholar] [CrossRef]
  49. Kumar, V.; Singh, J.K. Model for calculating the refractive index of different materials. Indian J. Pure Appl. Phys. 2010, 48, 571–574. [Google Scholar]
  50. Pesika, N.S.; Stebe, K.J.; Searson, P.C. Relationship between Absorbance Spectra and Particle Size Distributions for Quantum-Sized Nanocrystals. J. Phys. Chem. B 2003, 107, 10412–10415. [Google Scholar] [CrossRef]
  51. Goh, E.G.; Xu, X.; McCormick, P.G. Effect of particle size on the UV absorbance of zinc oxide nanoparticles. Scr. Mater. 2014, 78–79, 49–52. [Google Scholar] [CrossRef]
  52. Doak, J.; Gupta, R.K.; Manivannan, K.; Ghosh, K.; Kahol, P.K. Effect of particle size distributions on absorbance spectra of gold nanoparticles. Physica E 2010, 42, 1605–1609. [Google Scholar] [CrossRef]
  53. Gharibshahi, E.; Saion, E. Influence of Dose on Particle Size and Optical Properties of Colloidal Platinum Nanoparticles. Int. Mol. Sci. 2012, 13, 14723–14741. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Schematic representation of the preparation method of metallic M° and metal oxides MxOy nanoparticles inside M′xO′y matrices.
Figure 1. Schematic representation of the preparation method of metallic M° and metal oxides MxOy nanoparticles inside M′xO′y matrices.
Nanomaterials 10 02470 g001
Figure 2. (a) XRD pattern; (b) SEM image; (c) TEM image; (d) particle histogram; (e,f) HRTEM images; and (g) SEM element mapping of the pyrolytic NiO compound obtained using the chitosan precursor.
Figure 2. (a) XRD pattern; (b) SEM image; (c) TEM image; (d) particle histogram; (e,f) HRTEM images; and (g) SEM element mapping of the pyrolytic NiO compound obtained using the chitosan precursor.
Nanomaterials 10 02470 g002
Figure 3. (a) XRD pattern; (b) SEM image of NiO from chitosan and (c) from PVP; (d,e) TEM images of NiO from chitosan and (f) their histogram; and (g) SEM mapping element for NiO from the chitosan precursor.
Figure 3. (a) XRD pattern; (b) SEM image of NiO from chitosan and (c) from PVP; (d,e) TEM images of NiO from chitosan and (f) their histogram; and (g) SEM mapping element for NiO from the chitosan precursor.
Nanomaterials 10 02470 g003
Figure 4. (a) XRD pattern of NiO/Al2O3 from the chitosan precursor; (b) SEM image of NiO from chitosan and (c) from PVP; (d) TEM image of NiO from chitosan and (e) from PVP; (f) EDS mapping of NiO from chitosan.
Figure 4. (a) XRD pattern of NiO/Al2O3 from the chitosan precursor; (b) SEM image of NiO from chitosan and (c) from PVP; (d) TEM image of NiO from chitosan and (e) from PVP; (f) EDS mapping of NiO from chitosan.
Nanomaterials 10 02470 g004
Figure 5. (a) XRD pattern of NiO inside Na4.2Ca2.8 (Si6O18); (b) and (c) SEM images; and (d) EDS mapping by an element of the composite NiO/Na4.2Ca2.8(Si6O18).
Figure 5. (a) XRD pattern of NiO inside Na4.2Ca2.8 (Si6O18); (b) and (c) SEM images; and (d) EDS mapping by an element of the composite NiO/Na4.2Ca2.8(Si6O18).
Nanomaterials 10 02470 g005
Figure 6. (a) Normalized concentration changing of methylene blue (MB) without the catalyst and in the presence of NiO from the PVP and (b) their zero order kinetic of degradation of MB, (c) changing of MB without the catalyst and in the presence of NiO from the chitosan and (d) their zero order kinetic of MB degradation.
Figure 6. (a) Normalized concentration changing of methylene blue (MB) without the catalyst and in the presence of NiO from the PVP and (b) their zero order kinetic of degradation of MB, (c) changing of MB without the catalyst and in the presence of NiO from the chitosan and (d) their zero order kinetic of MB degradation.
Nanomaterials 10 02470 g006
Figure 7. Photocatalytic behavior of NiO embedded in several matrices.
Figure 7. Photocatalytic behavior of NiO embedded in several matrices.
Nanomaterials 10 02470 g007
Figure 8. Variation of both Eg and λmax with the different matrices.
Figure 8. Variation of both Eg and λmax with the different matrices.
Nanomaterials 10 02470 g008
Figure 9. Relationship between Eg and λmax for the composites the NiO/SiO2, NiO/TiO2, NiO/Al2O3, and NiO/Na4.2Ca2.8(Si6O18).
Figure 9. Relationship between Eg and λmax for the composites the NiO/SiO2, NiO/TiO2, NiO/Al2O3, and NiO/Na4.2Ca2.8(Si6O18).
Nanomaterials 10 02470 g009
Table 1. Composition of the pyrolytic products from the respective precursors.
Table 1. Composition of the pyrolytic products from the respective precursors.
PrecursorPrecursor FormulaMatrixCompositeComposite Number
(1)Chitosan·NiCl2 (chitosan)-NiOC1
(2)PSP-4-PVP·NiCl2 (PVP)-NiOC2
(3)Chitosan·NiCl2SiO2NiO/SiO2C3
(4)PSP-4-PVP·NiCl2SiO2NiO/SiO2C4
(5)Chitosan·NiCl2TiO2NiO/TiO2C5
(6)PSP-4-PVP·NiCl2TiO2NiO/TiO2C6
(7)Chitosan·NiCl2Al2O3NiO/Al2O3C7
(8)Chitosan·NiCl2Na4.2Ca2.8(Si6O18)NiO/Na4.2Ca2.8(Si6O18)C8
Table 2. Nanoparticle size for the composites.
Table 2. Nanoparticle size for the composites.
CompositePrecursor FormulaParticle Size (nm)Reference
NiOChitosan·NiCl2>50[17]
NiOPSP-4-PVP·NiCl2>50[17]
NiO/SiO2Chitosan·NiCl225This work
NiO/SiO2PSP-4-PVP·NiCl2100This work
NiO/TiO2Chitosan·NiCl225This work
NiO/TiO2PSP-4-PVP·NiCl263This work
NiO/Al2O3Chitosan·NiCl230This work
NiO/Al2O3Chitosan·NiCl217This work
NiO/Na4.2Ca2.8(Si6O18)Chitosan·NiCl2Not measuredThis work
Table 3. Kinetic data for the photodegradation process of MB with NiO and NiO/SiO2, NiO/TiO2, NiO/Al2O3, and NiO/Na4.2Ca2.8(Si6O18) composites.
Table 3. Kinetic data for the photodegradation process of MB with NiO and NiO/SiO2, NiO/TiO2, NiO/Al2O3, and NiO/Na4.2Ca2.8(Si6O18) composites.
PhotocatalystApparent Photodegradation *Discoloration Rate (%)R2 Linear Fit (%)
NiO-CHITOSAN 2.471%0.998
NiO-PS-4-PVP2.268%0.991
NiO/SiO2-CHITOSAN2.369%0.999
NiO/SiO2-PS-4-PVP1.648%0.996
NiO/TiO2 -CHITOSAN2.991%0.992
NiO/TiO2-PS-4-PVP2.681%0.980
NiO/Al2O3-CHITOSAN1.545%0.990
NiO/Na4.2Ca2.8(Si6O18)2.675%0.990
* Rate constant k (10−3 M·min−1).
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.; 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. https://doi.org/10.3390/nano10122470

AMA Style

Diaz C, Valenzuela ML, Cifuentes-Vaca O, Segovia M, Laguna-Bercero MA. 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(12):2470. https://doi.org/10.3390/nano10122470

Chicago/Turabian Style

Diaz, Carlos, María L. Valenzuela, Olga Cifuentes-Vaca, Marjorie Segovia, and Miguel A. Laguna-Bercero. 2020. "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 10, no. 12: 2470. https://doi.org/10.3390/nano10122470

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