Next Article in Journal
Phase Transitions in Boron Carbide
Next Article in Special Issue
Deep Red Photoluminescence from Cr3+ in Fluorine-Doped Lithium Aluminate Host Material
Previous Article in Journal
The Effect of Chemical and Thermal Treatment for Desizing on the Properties and Chemical Functional Groups of Carbon Fiber
Previous Article in Special Issue
Optical Properties Investigation of Upconverting K2Gd(PO4)(WO4):20%Yb3+,Tm3+ Phosphors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Additives in Nanocrystalline Tin Dioxide: Recent Progress in the Characterization of Materials for Gas Sensor Applications

Chemistry Department, Moscow State University, Moscow 119991, Russia
*
Author to whom correspondence should be addressed.
Materials 2023, 16(20), 6733; https://doi.org/10.3390/ma16206733
Submission received: 23 August 2023 / Revised: 12 October 2023 / Accepted: 13 October 2023 / Published: 17 October 2023

Abstract

:
Tin dioxide has huge potential and is widely studied and used in different fields, including as a sensitive material in semiconductor gas sensors. The specificity of the chemical activity of tin dioxide in its interaction with the gas phase is achieved via the immobilization of various modifiers on the SnO2 surface. The type of additive, its concentration, and the distribution between the surface and the volume of SnO2 crystallites have a significant effect on semiconductor gas sensor characteristics, namely sensitivity and selectivity. This review discusses the recent approaches to analyzing the composition of SnO2-based nanocomposites (the gross quantitative elemental composition, phase composition, surface composition, electronic state of additives, and mutual distribution of the components) and systematizes experimental data obtained using a set of analytical methods for studying the concentration of additives on the surface and in the volume of SnO2 nanocrystals. The benefits and drawbacks of new approaches to the high-accuracy analysis of SnO2-based nanocomposites by ICP MS and TXRF methods are discussed.

1. Introduction

Tin dioxide SnO2 is an important representative of the group of n-type wide-bandgap (Eg = 3.6 eV) semiconductor oxides (SnO2, ZnO, In2O3, TiO2, and WO3 [1]) with a unique combination of high transparency, electrical conductivity, high surface reactivity, and stability in air. Tin dioxide has huge potential and is widely studied and used for applications in different fields [2,3], in particular for the manufacturing of optically passive components in a number of devices [4,5,6], including photodetectors [7,8,9,10,11,12], solar cells [13,14,15,16], conductive ceramics [17,18,19,20,21,22], varistors [23], transparent electrodes [24,25,26,27,28], electrochromic coatings [29,30,31,32,33,34], transistors [35,36,37], components of fuel cells [38] and metal-ion batteries [39], catalysts [40], and chemical sensors [2,41,42,43,44]. A wide range of applications require the development of technologies for producing tin dioxide with significantly different properties: high resistance for heaters, high conductivity for transparent electrodes and contacts, and a large specific surface area with selective adsorption properties and reactivity combined with appropriate conductivity for chemical sensors. It should be noted that bare tin dioxide without special additives does not have the abovementioned properties and is rarely used. For example, the required level of SnO2 electrical conductivity is achieved by introducing electroactive additives into the bulk of tin dioxide (doping), among which Sb, Nb, and F are the most widespread. The pentavalent Sb and Nb, which are close to Sn4+ in ionic radius, exhibit donor properties and increase the concentration of free charge carriers when replacing Sn4+ cations in the SnO2 crystal lattice, as well as fluorine F anions when replacing oxygen O2− anions.
Considering the material requirements for semiconductor gas sensors, it should be noted that the stability of SnO2 in air is necessary to ensure long-term functioning at the elevated operating temperatures that are usually realized when detecting gases; a large specific surface area is needed to increase the proportion of material in direct contact with the gas phase; the reactivity of the surface determines the possibility of the adsorption of the target gas molecules and the occurrence of the reactions responsible for the sensor response formation; and a suitable level of electrical conductivity should make it possible to register the changes when interacting with gases, both electron donors and acceptors. Discussing the use of SnO2-based materials for gas sensor applications, it is worth noting that the phase composition, the chemical composition of the surface, the spatial distribution of the components, and the real structure of the materials are determined by the synthesis conditions. In turn, these characteristics determine the reactivity of the materials and the mechanism of their interaction with the gas phase; the concentration and mobility of the charge carriers; and, consequently, the sensor parameters, namely sensitivity, selectivity, and stability. The specificity of the chemical activity of tin dioxide in the interaction with the gas phase is achieved via the immobilization of various modifiers on the SnO2 surface: clusters or nanoparticles of precious metals and their oxides (Au, PtOx, PdOx, and RuOx) or transition metal oxides (La2O3, Fe2O3, Co3O4, NiO, CuO, etc.) [45,46,47,48,49,50,51]. Modifiers affect the concentration of adsorption centers and the mechanism of reactions on the SnO2 surface [51]. The specificity of adsorption and the reduction in the reaction temperature for modified SnO2 are the result of a synergistic effect of a number of factors: the formation of acidic or basic centers; the activation of chemisorbed molecules; and, in some cases, the spillover effect [46,51,52]. The conscious choice of the modifier can be based on fundamental ideas about the acid–base properties of the catalytic modifier quantified via the value of the optical basicity [53,54,55] or the Sanderson electronegativity scale [56,57] used in heterogeneous catalysis for analyzing the adsorption and reaction activity of materials [46].
Complex inhomogeneous nanocrystalline systems—nanocomposites formed by introducing additives into a highly dispersed tin dioxide matrix [45,46,47,48,49,50,51,52]—are systems constructed of nanometer-sized crystalline particles (3–30 nm) combined into agglomerates, in which the additive is distributed in a complex way between the volume and the surface of the crystalline grains of the main phase [58]. Considering that in nanocrystalline systems (which are non-equilibrium systems), surface atoms make a significant contribution to the interaction of the main phase and the modifier, it is impossible to predict the nature of the mutual distribution of components in such systems using equilibrium P-T-x phase diagrams. In Gibbs’s approach, interface phases (including the solid/gas interface) are modeled by the combination of a phase possessing an area but no volume (a dividing surface) with surface excess properties accounting for all the effects of the interface [59]. By mathematically uniquely placing the dividing surface, all thermodynamic variables of the surface are accounted for in a way that captures all the effects of the interface. This is an important concept to understand because it is not to say that the approach of Gibbs can only be applied to interfaces that have no reach into the adjoining bulk phases, but rather that the excess molecules and energy of that reach must be added to the presence of a uniquely defined infinitely thin surface to completely describe the system. In bulk systems, the surface effect is negligible, so the contribution of the surface thermodynamic functions to the thermodynamic description of the system as a whole (graphically represented by P-T-x phase diagrams) is absent. This makes equilibrium P-T-x phase diagrams inapplicable for describing highly dispersed systems, including nanocrystalline solids. For example, it has been shown that, unlike microcrystalline systems, the regions of the existence of solid solutions and the conditions of phase formation in nanocrystalline materials significantly depend on the dispersion of samples [45]. It was established in [45,48,60] that the introduction of additives leads to an increase in the thermal stability of nanocomposites compared with individual oxides. Due to the significant proportion of surface layers with a high degree of structural disordering, grain growth in nanocrystalline systems is accompanied by the structural relaxation of the surface. The release of an additive in the form of the segregation of an amorphous phase on the surface of growing SnO2 crystallites slows down the rate of their growth under isothermal annealing [45] due to the so-called Zener pinning of the matrix grain boundaries [61], according to which the maximum size of a crystalline grain is determined by the volume fraction and size of particles segregated on the surface of the growing crystallites. Depending on the ratio of components, nanocomposites also have different phase compositions. In addition to the formation of solid solutions, the formation of additive segregation on the surface of the crystalline grains of the main phase and the stabilization of thermodynamically unstable phases are possible [62].
Experimental studies on “composition–structure–property” relationships indicate that the additives that are used to achieve the necessary functional parameters of tin dioxide, depending on the concentration and local position in the bulk or on the surface, perform the function of electroactive dopants or electrical/chemical active modifiers. The effect that additives have on SnO2’s properties primarily depends on the element’s electronic structure, structural parameters, and acid/base and redox properties. Given that additives are distributed in a complex way between the surface and the bulk of SnO2 crystallites, the implementation of only one function is impossible. In turn, the additive’s distribution between the surface and the volume of SnO2 crystallites has a significant effect on the sensor’s characteristics, namely sensitivity and selectivity.
Depending on the electronic structure of the atoms, the additives that have been most used for SnO2 functionalization can be divided into the following groups:
  • p-elements—Sb and F;
  • 3d-, 4d-, and 5d-elements—Ti-Zn, Nb, Mo, and W;
  • Rare earth elements—La and Ce;
  • Precious metals and platinum group metals—Pt, Pd, Ru, Rh, Au, and Ag.
The complexity of interpreting the dependence of the functional properties on the nature and concentration of the introduced additive is caused by the fact that most of them contain elements that can exhibit different electronic states in the volume and on the surface of SnO2 crystallites [63,64,65,66,67]. Thus, the correct determination of the composition of the nanocomposite is necessary for the development of technology to obtain functional materials based on tin dioxide. It should be noted that in the case of nanocomposites, the term “composition” is complex and includes the gross quantitative elemental composition; phase composition; surface composition; electronic state of the additives and main components; and distribution of the additives between the volume and surface of the SnO2 crystallites.
Currently, there is no single technique devoted to the creation of materials for gas sensors that allows for the conscious choice of an additive in order to selectively increase the sensitivity of semiconductor oxides to a certain type of detected gas. The most promising, in our opinion, are catalytic approaches based on the concepts of “collective” and “local” centers. The consideration within the framework of the “collective” centers approach is based on the theory of chemisorption on the surface of semiconductors, developed in the works of T. Wolkenstein [68], which gives an idea of the mechanism of the adsorbate effect on the overall band structure of the modified semiconductor matrix. In this case, the catalytic activity of the modifier is directly related to the electronic state of the doping additives in the oxide matrix and their effect on the concentration of charge carriers in the semiconductor. On the other hand, the approach of “local” centers is based on the concept of an inhomogeneous surface. The interaction of the semiconductor oxide with the gas phase in this case is described through the formation of surface complexes, while the chemical nature of the modifier and its reactivity in acid/base or redox reactions play a decisive role. It is obvious that the regularities established in the study of catalysts may not be observed when applied to sensor materials and should be tested experimentally. This requires, among other things, a detailed characterization of semiconductor materials in order to reliably establish the patterns of “composition–structure–functional property” relationships.
This review discusses recent approaches to analyzing the composition of SnO2-based nanocomposites that act as sensitive materials for semiconductor gas sensors and systematizes the experimental data obtained by various authors using a set of modern analytical methods for studying the concentration and distribution of additives between the surface and volume of SnO2 crystallites. A comparison of the characteristics of the analytical methods discussed in this review is presented in Table 1. Diagrams illustrating the analytical dimension against the detection range and depth of analysis can be found in [69].

2. Gross Quantitative Elemental Composition

Despite the huge amount of experimental data on the effect of the concentration of additives on SnO2 sensor properties, a strategy for the quantitative elemental analysis of tin-dioxide-based nanocomposites has not been proposed. To characterize nanocrystalline SnO2, approaches that are convenient for analyzing the surface of solid samples and thin films are more often used; therefore, X-ray methods have become widespread. The chemical composition and quantitative ratio of components for nanoscale functional materials are most often determined by energy-dispersive X-ray spectroscopy (EDX) techniques in combination with electron microscopy. This approach is actively developing [70].
Energy-dispersive X-ray spectroscopy (EDX) is an analysis method that can be coupled with several applications, including SEM, TEM, and STEM. EDX, when combined with an electron microscope, can provide elemental analysis for areas as small as several nanometers. The effect of the electron beam on the surface and its penetration into the particles produces X-rays that are characteristic of the elements found on and near the surface of the sample. Using the EDX method, one can determine the elemental composition of individual points or map out the areal distribution of elements based on the scanning capability of the electron microscope. EDX does not fit into the category of techniques for surface science, as the X-rays are generated in a region about 2 microns in depth [69].
The accuracy of quantitative analysis by the EDX method is influenced by the porosity of the sample, its micro-uniformity (the presence of two or more phases), the material density, the peak overlap, the matrix composition, and the amount of additive. For analysis by the EDX method, it is necessary to choose a sample thickness that is, on the one hand, sufficient to ensure a good signal-to-noise ratio and, at the same time, thin enough to reduce the effects of electron scattering [71].
The advantage of X-ray spectra, in comparison with the spectra in the visible and ultraviolet regions, is the small number of lines for each element. When an atom with a vacancy passes from the excited to the ground state, an Auger electron can be emitted instead of X-ray characteristic photons. The probability of the emission of characteristic X-rays increases with an increasing atomic number, while the probability of the emission of Auger electrons decreases. Thus, the EDX method is most useful for the analysis of heavy elements, especially when the element concentration is low. The area of interaction between electrons and matter increases rapidly with the energy of the incident electron beam. The penetration depth decreases with an increasing material density. Dense metals, such as W and Au, have a limited electron transmission region, and, consequently, EDX spectra are more sensitive only for the surface layer of a sample of such materials. The results of quantitative X-ray analysis can be achieved with an accuracy of up to 1%. These results require careful analysis and should be obtained using standard samples. Most EDX computer programs work using quantification without standards, though this does not always lead to a reliable result. The most common method of quantification is known as the three-correction method. This method consists in measuring the X-ray signal I from an element in an unknown sample, and then the data obtained are compared with the peak signal for the same element in the reference sample, taking into account the correction coefficients: the atomic number, absorption, and fluorescence correction. It is important to note that any method of calculation assumes that the sample is flat and homogeneous in the volume excited by X-ray radiation. If X-rays are emitted from an area with a different composition or from an uneven surface, the absorption and fluorescence corrections may be inaccurate [72].
The geometry of the samples can also introduce significant uncertainty into the results of quantitative analysis. An interesting way to account for this factor by collecting information using several detectors arranged symmetrically relative to the sample was proposed by the authors of [73]. Some of the factors affecting the accuracy of the analysis can be taken into account using standard samples with the same surface conditions; however, there are no such samples for new materials. These difficulties can lead to an incorrect interpretation of the obtained dependencies of various material characteristics on the composition determined by the EDX method.
For example, for single-phase (according to XRD data) SnO2 samples doped with Fe by co-precipitation, the dependences of the cell parameters, volume of the unit cell, and sizes of the SnO2 crystallites on the concentration of added iron were shown in [74]. In this case, the sensitivity of the EDX method was sufficient to detect iron in the concentration range used, and the step of changing the concentration by 3–5 wt.% allowed the authors to correlate the structural characteristics with the iron content, but without evaluating the accuracy and locality of these results. It should also be noted that the results, obtained in this work by 57Fe Mössbauer spectroscopy, indicated the release of some Fe3+ cations into amorphous α-Fe2O3 on the SnO2 crystallites’ surface, but no estimates of the concentration of Fe3+ on the surface and in the volume of the SnO2 grains were given. The authors of [75], when characterizing SnO2 microtubules synthesized by the catalyst-free vapor–solid (VS) thermal method, discussed the differences in the results of the EDX analysis of the chromium content on the inner and outer surfaces of the tubes at the level of 1.0 and 1.2 at%, respectively, while there is no reason to believe that these results belong to different confidence intervals.
The reliability of determining the additive concentration at the level of 1–5 wt.% by the EDX method is often questionable. For example, in [76], to study the purity of the chemical composition of Co-doped SnO2 nanoparticles obtained via the hydrothermal method, samples were analyzed using EDX. In this case, the cobalt content in the material at a level of less than 1 wt.%, determined by the internal normalization method, was only an approximate estimate. In [77], 0.5, 1, 2, and 3 wt.% Pd-loaded 6 mol% Sm-doped SnO2 nanoparticles prepared using a co-precipitation method were investigated. For a composite with a pre-assigned Pd content of 3 wt.%, a 2.13 wt.% Pd content was found by the EDX method, but the accuracy of the results to hundredths of a percent was not confirmed. With an increase in the concentration of the additive, the EDX signal could be reliably detected, and the results obtained could be used to assess the composition of the samples. Given that the accuracy of the EDX determination of light elements, in particular oxygen, is low, relative ratios are often used to determine the concentration of additives in oxide systems, which can be estimated by the fundamental parameter method [78,79]. There are very few data verifying the correctness of the results of composition analysis by the EDX method. Thus, the authors of [80], for SnO2-Ag nanocomposites obtained by the sol–gel method, confirmed the results of EDX determination (3.7 wt.% Ag) by ICP-MS (4.1 wt.% Ag), while the method of sample dissolution was not specified. In [81], for a correct analysis of the elemental composition of SnO2/Fe2O3 nanocomposites by the EDX method, ceramic samples with different [Fe]/[Sn] ratios prepared from appropriate mechanical mixtures of oxides were used as reference materials.

3. Phase Composition

The analysis of the phase composition of nanocrystalline materials is a difficult task because of the significant broadening of the diffraction maxima; furthermore, traditional X-ray diffraction studies require the use of additional methods.
In SnO2-based nanocomposites, the crystalline phase containing the additive is often detected only at a sufficiently high concentration of the introduced additive and high annealing temperatures; in other cases, it is possible to form a solid solution and/or segregation of the amorphous phase containing the additive on the surface of the SnO2 crystallites. The implementation of one or the other option is determined by both fundamental reasons—for example, the difference in the ionic radii of Sn4+ (0.69 Å) and the additive’s cation Mn+—and the synthesis conditions—the method of introducing the additive, the concentration of the additive, and the temperature treatment. Based on the values of the ionic radii of metal cations in an oxygen octahedral environment [82], it can be assumed that the formation of SnO2-based solid solutions is possible upon the introduction of Sb5+ (0.60 Å) [83,84,85,86]; Ti4+ (0.605 Å) [87,88,89]; Cr3+ (0.615 Å) [65,90]; Mn3+ (0.645 Å, high spin) [64,91,92,93]; Fe3+ (0.645 Å, high spin) [81,94,95,96,97]; Co2+ (0.65 Å, low spin) [63,95,96,98,99,100,101]; Ni2+ (0.69 Å) [95,96,102,103,104,105,106]; Cu2+ (0.73 Å) [107,108,109]; Zn2+ (0.745 Å) [110,111,112]; Nb5+ (0.64 Å) [113,114,115]; and Ru3+ (0.68 Å) [67,116,117,118,119,120]. However, it is unlikely in the case of Pd2+ (0.86 Å) [121,122,123,124,125,126]; Pt2+ (0.80 Å) [127,128]; La3+ (1.06 Å) [129,130]; and Ce4+ (0.80 Å) [131,132,133]. An unambiguous conclusion about the formation of a solid solution can be made if there is a change in the parameters of the SnO2 crystal lattice with an increase in the concentration of the additive. However, in nanocrystalline systems, the exact determination of cell parameters from X-ray data is often difficult because of the significant broadening of the diffraction maxima. For this reason, valuable information about the phase composition of SnO2-based nanocomposites can additionally be obtained by transmission electron microscopy, electron diffraction, Raman spectroscopy, and Mössbauer spectroscopy [45,62,63,64,81,94,134,135].
Based on a large set of experimental data, a scheme describing the transformation of the phase composition of tin-dioxide-based nanocomposites with an increase in the concentration of the introduced additive was proposed in the review [45] (Figure 1). This idea was later extended to metal oxide nanocomposites formed by two metals, MI and MII, in the review [48]. With an increase in the additive content in SnO2-based nanocomposites, the formation of a solid solution based on tin dioxide, the segregation of the oxide of the second component in the form of a monolayer or islands on the surface of the crystallites of the main phase, and then the transition to a two-phase region were consistently observed. The positions of the boundaries x1, x2, x3, and x4 (Figure 1) are determined by the ratio of the ionic radii of Mn+ and Sn4+ and the defectiveness of the SnO2 structure. The extent of the areas of solid solutions is units of percents. Determining the position of the x1 and x4 boundaries is an extremely difficult task and requires the development of special analytical procedures. With an increase in the annealing temperature, the defectiveness of the SnO2 structure decreases, which leads to a decrease in the solubility of the second component.
The synthesis method influences the distribution of the additive between the volume and surface of the crystalline grains of SnO2 due to the nonequilibrium of nanocrystalline systems. Thus, the introduction of an additive into a preformed SnO2 matrix (impregnation, layer deposition, etc.) leads to its accumulation on the surface of the SnO2 crystallites (Table 2). This leads to the supersaturation of the additive’s precursor on the surface of the SnO2 matrix and provides the possibility of forming the phase of the second component through heterogeneous nucleation. Since diffusion in the solid phase is slow, kinetic difficulties do not allow the second component to fully penetrate into the crystal structure of the main phase. This is an additional parameter that facilitates the crystallization of the second component in the form of its own oxide phase. The use of methods involving the synthesis of multicomponent nanocrystalline materials from a common source containing tin and additive precursors (co-deposition from the liquid phase, hydrothermal synthesis, the pyrolysis of aerosols, flame spray pyrolysis, etc.) (Table 3) provides a high degree of component homogenization, short diffusion paths, and the absence of conditions for the supersaturation and nucleation of the phase of the second component. The combination of these factors benefits solid solution formation.
The effect of the synthesis method on the phase composition of SnO2/NiO materials can be illustrated by comparing the results obtained in [102,158] for SnO2−NiO core−shell nanowires produced in a two-step process (single crystalline SnO2-core nanowires were synthesized by vapor−liquid−solid deposition and then decorated with a polycrystalline NiO-shell layer by atomic layer deposition) [158] and those obtained for Ni-doped SnO2 nanorods produced by one-pot solvothermal synthesis in [102]. In the first case, a conformal coating of a polycrystalline NiO-shell layer (2–8.2 nm thickness) onto single crystalline SnO2 nanowires was confirmed by grazing incidence X-ray diffraction (GIXRD) and high-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) with EDX mapping. The authors of [158] did not discuss the possibility of the formation of solid solutions near the SnO2/NiO boundary. On the contrary, one-pot solvothermal synthesis allowed the authors of [102] to obtain Ni-doped porous SnO2 materials. For samples with the composition set during the synthesis as [Ni]/[Sn] = 0, 5, 10, and 15 mol.%, the unit cell parameters were analyzed by Rietveld refinement. It was found that when the [Ni]/[Sn] ratio was set at less than 10 mol.%, the unit cell parameters decreased with the increase in the Ni2+ concentration. However, for the sample with [Ni]/[Sn] = 15 mol.%, the unit cell parameter increased. The authors attributed such a change in the unit cell parameters to the achievement of the Ni2+ solubility limit in the form of substitution defects in the SnO2 structure. Unfortunately, there was no information in the article about the Ni2+ concentration in the SnO2 bulk; however, it was indicated that the actual Ni content in the samples determined by the XPS method varied from 1.8 to 6.3 mol.%. Given that XPS has a small depth of analysis (Table 1), the values obtained may have been overestimated due to nickel segregation on the surface of the SnO2 crystalline grains.
Quantitative XRD phase analysis [170] is based on comparing the relative intensities of diffraction maxima using the Reference Intensity Ratio (RIR)—the ratio of the absolute intensities of the most intense reflections of a given substance and corundum (α-Al2O3) on a diffractogram of a mixture containing 50 wt.% of components. The sensitivity of the analysis varies for different phases, and the minimum detectable concentration is 5–15 wt.%. This approach is rarely used [81,171] and has no practical application for the analysis of nanocrystalline materials since it requires additional calibration using other analysis methods.

4. Surface Composition and Electronic State of Additives

Information about the valence states of the elements, the chemical composition of the surface, and the interfaces that determine the properties of nanocrystalline materials can be obtained by X-ray photoelectron spectroscopy (XPS) [172,173]. The XPS method allows one to detect all elements except hydrogen. The detection limit is usually 0.1–1 at.%, and the accuracy is 0.3–1 at.%. XPS is usually not considered a method that provides a lateral resolution in the nanometer range; however, the depth of analysis is determined by the electron free path relative to inelastic collisions and is 0.5–2.5 nm for metals and 4–10 nm for organic substances. Conventional XPS operates at vacuum levels of about 10−9 mbar, since the signal intensity decreases because of the higher electron scattering losses within the gas or liquid medium.
The study of the fine structure of the O1s oxygen spectrum on the surface of nanocrystalline materials based on SnO2 allows one to evaluate the effect of the introduced modifiers on the number of chemisorbed oxygen-containing particles that can act as the Lewis or Brønsted base when interacting with gas-phase molecules [51]. The asymmetric O1s signal of nanocrystalline SnO2-based materials can usually be simulated by two components: the major signal with a binding energy of 530.0–530.5 eV due to bulk O2− anions and the lower peak of surface oxygen-containing species (various forms of chemisorbed oxygen and surface hydroxogroups) at higher a binding energy (531.0–533.0 eV). Determining the ratio between the intensities of the spectrum components corresponding to bulk O2− anions and surface oxygen species allows one to draw conclusions about the effect of additives on the concentration of the basic and oxidizing centers on the surface of SnO2-based materials and establish appropriate relationships with their sensor properties when detecting gases of various natures [42,45,46,51,63,64].
The influence of the predominant form of chemisorbed oxygen on the sensitivity of semiconductor sensors based on nanocrystalline SnO2 was theoretically justified in [174]. This effect is expressed in a change in sensitivity—the slope of the dependence of the sensor response on the partial pressure of the detected reducing gas, drawn in double logarithmic coordinates. Experimentally, the effect of the additive on the predominant form of chemisorbed oxygen and sensitivity when detecting CO and H2 was established for SnO2/SiO2 materials [175,176,177]. For SnO2/SiO2 core–shell nanofibers obtained via template synthesis, under the introduction of water vapor at 400 °C, the predominant oxygen species remained both O2− and O. Meanwhile, under the same conditions, the dominant oxygen species on the SnO2 nanoparticles were changed from O2− to O, which was accompanied by a decrease in sensitivity when detecting H2 [175]. For SnO2/SiO2 nanocomposites synthesized by the hydrothermal method, an increase in the silicon content led to an increase in the contribution of molecular O2 ions that was consistent with the data obtained by EPR spectroscopy [176]. This change in the predominant form of chemisorbed oxygen was accompanied by an increase in sensitivity and a decrease in the CO detection limit, including in humid air [177].
The results of the quantitative analysis of SnO2-based nanocrystalline material composition by the XPS method are rarely discussed. Thus, the ratio of photoelectronic line intensities was used to approximate the ratio of Ru and Sb on the surface of SnO2 films in [178]. The authors of [179] obtained results for the determination of the Sb content on the surface of SnO2 synthesized by the vapor–liquid–solid process, while the inclusion of antimony in the SnO2 lattice was confirmed by Raman spectroscopy. The authors presented the antimony concentration with an accuracy of hundredths of a percent (8.19 at.%) but did not indicate the error of determination, and the relationship between the obtained quantitative results and the properties of the semiconductor nanowires was not discussed.
When analyzing the active centers created in gas-sensitive SnO2-based nanocrystalline materials, it is important to determine the oxidative state of the introduced additive. For the noble metals used, the tendency to form oxides decreases in the raw: Ru, Rh, Pd, Pt, Au. A combination of XPS and EPR methods showed that ruthenium in SnO2/RuOy nanocomposites (1 wt.% Ru), obtained by SnO2 impregnation with Ru(acac)3 solution and subsequent thermal annealing, could be in a mixed-valence state due to the formation of RuO2 clusters, including a fraction of Ru3+ [67,116]. Gold, even in the form of small clusters (1–3 nm) on the surface of sensor materials, is usually present in the zero–valent state [180,181,182]. However, components corresponding to Au3+ were found in the Au4f XP spectra of SnO2/SiO2–Au nanocomposites, which was explained by the ability of Au3+ (0.68 Å) to be embedded in the positions of coordination-unsaturated tin cations present on the partially reduced surface of SnO2 [183]. An additional effect exerted by SiO2 nanoparticles is the stabilization of Au cationic forms by hydroxyl groups due to the formation of Auδ+−OH bonds [184], with silanol groups making a significant contribution to the formation of these bonds.
Palladium and platinum can both be in the metallic state and in the form of PdO, PtO, and PtO2 oxides. The oxidation degree in this case depends on the nanocomposites’ synthesis procedure, the morphology and size of the additive particles, and the temperature and composition of the surrounding gas phase. XPS studies of the state of palladium deposited on SnO2 have shown that in low concentrations (0.1–0.2 wt.%), the modifier is distributed over the surface at the atomic level, which facilitates its interaction with chemisorbed oxygen to form PdO [121,122]. With an increase in the content to 1–3 wt.%, palladium forms three-dimensional clusters on the SnO2 surface [121], which in the presence of oxygen already begin to oxidize at 200 °C. As a result, mixed-valence PdOx clusters are formed containing Pd2+ as the main form (x = 0.7—0.8) along with zero-valent palladium [122,123,124]. Pd2+ cations can be embedded in the SnO2 lattice at the Sn4+ positions [121]; however, heating above 400 °C is required to start Pd2+ diffusion into the near-surface layer of the SnO2 matrix [125]. A detailed study of the composition of SnO2/PdOx (1 wt.% Pd) nanocomposites by a complex of methods (XPS, EPR, XANES, and EXAFS) revealed that PdOx clusters consisting of amorphous PdO also included Pd0 atoms and a small fraction of Pd3+ cations, presumably stabilized at the boundary of the clusters and the SnO2 matrix [67,116]. A further increase in the amount of palladium deposited led to the percolation of metal clusters; the formation of continuous Pd layers; and, finally, a near-surface solid solution of Sn–Pd (2–5 monolayer additives) [121]. The XPS studies carried out by the authors of [126] showed that in SnOx-decorated Pd nanocubes (the estimated edge length of the cubes was 21 ± 7 nm), palladium was mainly present as Pd0, but a small amount of PdO and PdO2 was also detected. The XPS results did not indicate the intermixing of Pd and SnO2 and the formation of an interfacial PdSn alloy.
In a similar way, the particle size affects the state of platinum on the surface of nanocrystalline SnO2: it has been shown that clusters smaller than 3–4 nm represent PtO2, and larger ones represent Pt0 [127]. According to XPS, in SnO2/PtOx (1 wt.% Pt) nanocomposites obtained by SnO2 impregnation with Pt(acac)2 and subsequent thermal annealing, platinum presents as PtO, which is typical for small (less than 2 nm) Pt clusters. In the case of a SnO2/PtPdOx nanocomposite containing bimetallic PdPtOx clusters, the Pd3d XP spectrum was shifted toward lower energies, indicating the partial reduction of Pd, while the Pt4f XP spectrum corresponded to the Pt2+ oxidation state in PtO [128].
The unambiguous determination of the electronic state of 3D elements by the XPS method can be difficult due to the similar electron binding energies for different electronic states [185]. For example, in the Ti2p XP spectra of hollow TiO2-SnO2-TiO2 composite nanofibers (the body consisted of hollow SnO2 nanofibers, fabricated by a facile electrospinning technique, and the inner and outer surfaces were covered by mixed-phase TiO2 nanoparticles grown via the hydrothermal method) [186], the Ti 2p3/2 peak was broad and asymmetric, demonstrating that there were at least two kinds of Ti chemical states. After curve fitting, besides the peak corresponding to Ti4+, another peak with a lower binding energy was identified, suggesting the existence of Ti3+ (or Ti–SnO bonds at the interface between the TiO2 nanoparticles and SnO2 nanoparticles) in the composite.
In SnO2/MnOx nanocomposites ([Mn]/[Sn] = 1 at.%) obtained by SnO2 impregnation with Mn(acac)3 and subsequent thermal annealing [64], the manganese oxidation state could not be unambiguously determined from the Mn2p XP spectra, since the difference between the electron bond energy of Mn 2p3/2 peaks for Mn3+ and Mn4+ is only 0.5–1 eV [187,188]. Additional information could be obtained by Raman spectroscopy.
The Co2p XP spectra of SnO2/CoOx nanocomposites obtained by SnO2 impregnation with Co(NO3)2 and subsequent thermal annealing, depending on the total cobalt content in the nanocomposites and the annealing temperature of the SnO2 matrix, could be described by two or four doublets corresponding to Co3+ and Co2+ and satellite peaks of Co2+. In nanocomposites with a low annealing temperature (300 °C) for the SnO2 matrix, cobalt presented in the Co2+ oxidation state. In nanocomposites with a high annealing temperature (750 °C) for the SnO2 matrix with a total cobalt content of 1.5 at.%, only Co2+ was found, and an increase in the cobalt content led to the appearance of the Co3+ state [63]. Additional information obtained from EPR data made it possible to determine the fraction of paramagnetic Co2+ centers from the total cobalt content. Similarly, in Co-doped SnO2 nanoparticles (with a [Co]/[Sn] ratio of about 2 at.%) obtained via the hydrothermal method [76], cobalt presented only in the Co2+ state. In [101], cobalt-doped 3D inverse-opal SnO2 multilayer films were prepared using the ultrasonic nebulizing deposition (UND) method combined with a self-assembly template. The Co2p XP spectra were described using the Co3+ and Co2+ components, and the [Co3+]/[Co2+] ratio decreased with an increased total cobalt concentration. For cobalt-doped SnO2 nanobelts prepared by the chemical vapor deposition method [189] (the total cobalt content was not specified), the Co2p XP spectra did not allow us to make an unambiguous conclusion about the electronic state of the cobalt; however, based on a combination of results obtained by XRD and Raman spectroscopy, the authors of [189] claimed that Co2+ and Co3+ were successfully substituted into the SnO2 structure at the Sn4+ site. Note that this statement is controversial because of the significant difference in the ionic radii of Co2+ and Co3+.
Establishing the electronic state of 4f elements is an even more difficult task due to the complexity of XP spectra. However, the authors of [131] were able to identify certain patterns in the electronic state of cerium in CeO2-SnO2 nanocomposites. The results showed that the Ce3+ content in CeO2 was substantially lower than that of the CeO2-SnO2 composites, but the Ce3+ content was almost independent of the total Ce content. This indicates that the oxidation states of Ce are markedly affected by the presence of surrounding Sn species.

5. Distribution of Additives between the Volume and the Surface of SnO2 Crystallites

The distribution of the additive between the volume and the surface of the crystalline grains of the main phase plays an extremely important role in the formation of the functional properties of nanocomposites. The introduction of impurity atoms into the SnO2 crystal structure causes the formation of impurity levels, the compensation of donor oxygen vacancies by acceptor impurity defects, and the modulation of the band relief of the semiconductor. The segregation of the phase formed by the additive on the surface of SnO2 crystallites can lead to the formation of pn heterojunctions in the area of intergranular contacts, which also inevitably affects the electrophysical properties of the material. In addition, the distribution of impurities between the volume and the surface of the SnO2 crystalline grains determines its effectiveness in the interaction between the material and the gas phase. The combination of these factors has a significant impact on the sensor properties of nanocomposites. Thus, the search for approaches to the correct highly sensitive determination of the additive concentration in the volume and on the surface of the SnO2 crystalline grains is an urgent task.
The formation of clusters or nanoparticles of noble metals or their oxides localized on the surface of SnO2 has been experimentally confirmed by the study of nanocomposites via transmission electron microscopy [190,191,192,193,194]. It is possible to detect clusters if the atomic masses of the additive and the main element of the oxide matrix differ significantly (for example, when tin dioxide is modified with gold or nickel), or if the sizes of the modifier clusters differ significantly from the SnO2 particle size. On the contrary, when noble metals of the same period (Ru, Rh, Pd) are immobilized on the surface of nanocrystalline SnO2 (particle size less than 10 nm), the visualization of additives by transmission microscopy methods is difficult [124,158]. When studying the distribution of additives in such systems, it is informative to scan the elemental composition of the surface using energy-dispersive X-ray spectral microanalysis (EDX mapping). For SnO2/PdOx and SnO2/RuOx nanocomposites, it has thus been shown that these additives form clusters of 1–5 nm even on highly dispersed SnO2 (particle size 3–6 nm) [124,190,193]. In general, the combination of transmission microscopy methods with EDX mapping makes it possible to visualize the distribution of components in nanocrystalline materials, as well as to analyze the phase composition with high locality [124,131,137,158,171,186,195]. However, the insufficient sensitivity of the EDX method often does not allow one to reliably determine the concentration of the additive embedded in the SnO2 crystal structure.
The solubility of antimony in nanocrystalline tin dioxide is estimated as 3–6 at.% based on the dependence of the conductivity on the Sb concentration [195,196,197,198]. Diffraction analysis methods are ineffective for studying antimony distribution in nanocrystalline SnO2. Due to the similar size of Sb5+ and Sn4+ (0.65 Å and 0.69 Å, respectively), cation substitution does not affect the unit cell parameters. Because of the proximity of the atomic masses of Sn and Sb, it is not possible to visualize the impurity in SnO2(Sb) by contrast using electron microscopy methods. According to Mössbauer spectroscopy, EXAFS, and XANES data, at low concentrations, antimony is embedded in the SnO2 structure in the Sb5+ form, and when the doping level increases, it comes to the surface in the Sb3+ form [66].
Fabbri et al. used a comparison of the results obtained by the EDX and XPS methods to estimate the distribution of the Sb additive between the volume and the surface of SnO2 nanocrystals based on different signal acquisition depths [199]. However, it remains unclear what factors were taken into account when calculating the Sb content with an accuracy of tenths of a percent. Zaytsev et al. used Auger electron spectroscopy to study the distribution of antimony between the surface and volume of SnO2 whiskers synthesized by the vapor–liquid–crystal method [200]. Auger spectra were obtained on the surface and after the ion beam etching of a 100 nm layer on the same whisker. Antimony was clearly visible on the surface, and after etching, no antimony signal was detected in the spectrum; thus, it was shown that antimony was concentrated in the near-surface layer of the whiskers (the layer thickness did not exceed 100 nm). At the same time, the antimony content in the volume of the whiskers was below the detection limit of the Auger spectroscopy method, and the antimony content found on the surface significantly exceeded its total amount determined by laser-induced mass spectrometry.
The authors of [201] proposed for the first time the use of the highly sensitive ICP MS method for the determination of antimony in SnO2 whiskers, representing a new approach to studying the distribution of the additive between the surface and volume. The approach included the treatment of the whisker surface with citric acid to transfer Sb3+ from the surface to stable soluble complexes, the complete dissolution of the sample residue after flushing from the surface, and the dissolution of the initial sample followed by the analysis of all solutions using the ICP MS method. The uniqueness of the proposed approach was based on SnO2’s properties: the material is barely soluble, so it was possible to select a reagent for flushing the additive from the SnO2 surface without affecting the tin dioxide (Table 4). It was shown that the correct determination of the element concentrations could be carried out using standard aqueous solutions. It was found that the surface antimony concentration was (1.6 ± 0.3) 10−2 wt.%, while the total antimony content was (2.8 ± 0.3) 10−2 wt.%. The application of the proposed analysis technique made it possible to determine the composition of semiconductor materials with a standard deviation of sr = 0.10. Using this approach to analyze the composition of nanocrystalline SnO2 with Au and Sb additives, it was possible to establish that gold was immobilized on the SnO2 surface, while antimony was distributed between the surface and the volume of the SnO2 grains [202].
A similar approach using the partial and complete dissolution of nanocomposites and the subsequent analysis of the obtained solutions by the ICP MS method was employed to determine the concentration of chromium with sr = 0.07 and to study its distribution between the volume and the surface of SnO2 nanocrystals [90]. In an octahedral oxygen environment, the ionic radius of Cr3+ is smaller than that of Sn4+ (0.615 Å and 0.69 Å, respectively), which allows the chromium ion to be embedded in the crystal lattice of tin dioxide [203]. To determine the chromium concentration on the surface of SnO2 nanocrystals, the samples were treated with potassium rhodanide, after which the obtained solutions containing stable chromium rhodanide complexes were analyzed. The novelty of this approach should be noted, since chromium oxides are barely soluble and are effectively dissolved by fusing them with sodium carbonate and subsequent leaching with an alkali [204]. The chromium content in the crystal structure of SnO2 was determined from a solution obtained after the decomposition of the sample residue in a mixture of HCl + HF + HNO3 acids in a microwave autoclave. It was shown that, regardless of the amount of modifier introduced during the synthesis, in samples synthesized by the impregnation of the SnO2 matrix, 20–30% of the total amount of introduced chromium was on the surface of the materials, and 70–80% was distributed in the crystal structure of the SnO2. Using this approach, it was also shown that there was no chromium on the surface of materials synthesized by flame spray pyrolysis, which provides a high degree of mixing of precursors under synthesis conditions. Thus, this method of producing modified tin dioxide ensures the distribution of the additive only in the SnO2 crystal structure. The total amount of chromium in the samples was no more than 0.01 wt.%, which was an insignificant portion (1–3%) of the chromium introduced during synthesis. By this fact, the authors explained the deterioration of the sensor properties of materials synthesized by flame spray pyrolysis compared with materials obtained via impregnation.
A study of the distribution of elements in SnO2/CoOx nanocomposites carried out by the HAADF-STEM method with EDX mapping showed that in nanocomposites obtained at a low annealing temperature (300 °C) for the SnO2 matrix, cobalt was evenly distributed over the surface of the SnO2 crystallites, without forming particles of individual cobalt-containing oxide phases [63]. On the contrary, in nanocomposites synthesized using a high annealing temperature (750 °C) for the SnO2 matrix, there were both particles containing only tin and particles containing only cobalt. The quantitative analysis of the composition of nanocomposites was carried out by the ICP MS method. To assess the distribution of cobalt between the volume and the surface of SnO2 crystalline grains, sample preparation techniques have been developed that allow one to independently determine the cobalt content on the surface of SnO2 (without destroying the tin dioxide matrix) and the total cobalt content in nanocomposites (under conditions of the complete dissolution of the SnO2 matrix). In nanocomposites annealed at 300 °C, the amount of cobalt found on the surface of SnO2 grains does not exceed 30% of its total content in nanocomposites. This allows us to assume that in these samples, a portion of the cobalt is embedded in the SnO2 crystal structure. In these nanocomposites, cobalt plays the role of electroactive additive. For samples annealed at 750 °C, the amount of cobalt detected on the surface coincides with the total cobalt content within the error of the analysis, i.e., under these synthesis conditions, cobalt does not enter the tin dioxide crystal structure and is localized as cobalt-containing oxide phases on the surface of SnO2 crystalline grains. The formation of p-CoOx/n-SnO2 heterojunctions is responsible for the electrical properties of these nanocomposites and determines their sensor properties when detecting CO, NH3, and H2S.
To analyze the distribution of components in SnO2/CoOx/Au nanocomposites, in which Au is located on the SnO2 surface and cobalt can be distributed between the surface and the volume of the SnO2 crystallites [98], a new procedure for determining the concentration of additives separately in the volume and on the surface of SnO2 by the ICP MS method has been developed. The essence of the method consists in the preliminary dissolution of gold and cobalt oxide on the surface with a mixture of acids (HNO3:HCl = 1:3), centrifugation, and the decantation of the filler solution. The use of this method practically eliminates the transition of tin (which is the main component) into the solution. After dissolving the additives localized on the SnO2 surface, the samples are decomposed in a mixture of HNO3, HCl, and HF in a microwave oven, followed by the determination of the concentration of Co, Au, and Sn by the ICP MS method [205]. It should be noted that the determination of gold in the content range of 0.03–0.1 wt.% is possible only with the use of the highly sensitive ICP MS method.
A distinctive feature of the considered approaches is the need to transfer difficult-to-dissolve cassiterite into a solution suitable for ICP MS analysis. The successful resolution of this task using SnO2 cementation with zinc powder in hydrochloric acid was proposed in [201]. However, such a sample preparation technique is time-consuming and leads to the zinc contamination of the analyzed solutions. The approach of sample decomposition in an autoclave with microwave intensification seems to be more promising. The decomposition of samples in autoclaves within a microwave oven has a number of advantages: the ability to control the temperature and heating modes and no need to use large volumes of reagents, which led to a decrease in the signal of the blank experiment [138,206]. This method was used in [63,64,90,201] for the dissolution of various SnO2-based samples. At the same time, the acid mixtures selected for dissolution depend on the type of additives, and the furnace heating programs differ for SnO2 samples synthesized at different temperatures. With an increase in the synthesis temperature, the concentration of defects in the nanocrystals decreases, and particles become larger. Therefore, for samples synthesized at temperatures above 600 °C, a specific autoclave heating program is necessary. For example, the complete decomposition of SnO2/MnOx nanocomposites requires heating up to 220 °C in two stages for three hours [64]. In other cases, processing in a mixture of HCl + HF (+HNO3 for platinum group metals) for 30 min at 180 °C is sufficient to dissolve cassiterite [128].
The results of the separate determination of platinum and palladium concentrations using the previously proposed ICP MS approach were used to compare the synthesis methods of modified nanocrystalline SnO2 [207]. The synthesis of the SnO2 matrix was carried out by two methods: chemical precipitation and flame spray pyrolysis. Noble metals were introduced by the impregnation of the SnO2 matrix with subsequent thermal decomposition. It was established that palladium was located on the surface of the nanocomposites regardless of the synthesis procedure. At the same time, no loss of the additive was noted. It was shown that for SnO2 obtained by chemical precipitation, 30–50% of the Pt was distributed on the surface, and the remaining 50–70% was in the volume of the SnO2 grains. The use of flame spray pyrolysis for SnO2 synthesis increased the Pt content on the SnO2 surface by up to 80%. Platinum losses were noted for both synthesis methods, which may have been due to various factors. Based on the low reproducibility of the results of the Pt and Pd determination (sr = 0.35), the heterogeneity of the additives’ distribution in the studied materials was established.
Note that despite the developed procedure for dissolving SnO2 for ICP MS analysis, this stage of sample preparation increases the time and complexity of the analysis and can serve as a source of errors. An alternative may be the use of total reflection X-ray fluorescence (TXRF). An important feature of the TXRF method is the formation of standing waves on the surface under conditions of the complete external reflection of X-rays, which contributes to the effective excitation of the X-ray fluorescence of the sample under study. Another feature concerns the registration of characteristic radiation. In TXRF, the detector is placed very close to the surface of the sample holder, providing a large solid angle of radiation collection; effective registration; and, as a result, a high counting speed that reduces the analysis time to 100–1000 s. Due to the use of a thin layer of the sample, there is no phenomenon of X-ray reabsorption. Thus, when the thickness of the sample layer decreases below a certain critical value, matrix effects practically do not appear. In analytical practice, TXRF makes it possible to eliminate the typical disadvantages of energy-dispersive X-ray fluorescence analysis, such as low sensitivity, the presence of matrix effects, and the need for external calibration.
TXRF uses a unified geometry of 0.1°/90° (usually referred to as 0°/90°), and since samples are used for analysis in the form of thin layers, X-ray radiation is weakly scattered. This makes it possible to significantly reduce the level of background radiation and significantly improve the detection limits, reaching in some cases levels of μg/kg (or μg/L). In TXRF, the calculation is usually implemented according to the method of the internal standard, which eliminates the need for the preparation of standard reference samples. This provides great advantages in cases where the preparation of standard samples is difficult, for example, new nanocrystalline materials. In this way, ruthenium in SnO2 was determined at a concentration level below 1 wt.% using gallium as an internal standard. At the same time, it was shown that the reproducibility of the results of Ru determination by TXRF was higher when using solutions obtained via sample decomposition than in suspensions without sample preparation (sr = 0.09 and 0.15, respectively), because of the complex morphology of the surface formed after the suspension dried on the reflector substrate. In combination with the low matrix influence, this makes the TXRF method suitable for the analysis of complex objects [208,209,210]. Thus, the authors of [211] developed a method for determining Ru in SnO2 suspensions by the TXRF method without sample preparation, while the correctness of the determination results was confirmed by the ICP MS method after the analysis of sample solutions following decomposition. The authors of [211] also noted the limitations of the TXRF method. The definition of Pd was impossible, as the analytical line for Pd Lα (2.84 keV) partially intersected the line for Ar Kα (2.96 keV) from the air layer between the sample and the X-ray radiation detector. For this reason, the distribution of ruthenium and palladium between the surface and the volume of SnO2 nanocrystals was studied by the ICP MS method in solutions obtained after processing samples with concentrated HCl. It was shown that tin did not pass into solution together with Ru and Pd under the selected conditions. It was found that the concentration of Ru and Pd on the SnO2 surface was 0.4 and 0.2 wt.%, respectively. It was noted that it was optimal to use the ICP MS method to determine the noble metal additives at a level of less than 1 wt.% in samples of low weight.
According to [180,181], gold clusters are distributed on the surface of SnO2; therefore, the use of the TXRF method for sample suspensions significantly simplifies the analysis of the Au concentration in SnO2. In this case, it is possible to directly determine the total Au content at the level of 1–2 wt.% using gallium as an internal standard with sr = 0.12 [212]. Interestingly, the results of the analysis of suspensions by the TXRF method, in contrast to ICP MS, can be used to assess the uniformity of the additive distribution over the surface by evaluating the results of determination in individual aliquots taken from the volume of suspensions [213]. For the correct determination of Au at a concentration level of 0.2 wt.%, the sensitivity of the TXRF method is not sufficient due to the need for low weight sampling to meet the conditions of a thin layer when analyzing the sample. Therefore, it is proposed to use another approach to determine such low concentrations.
For SnO2-based nanocrystalline materials, a procedure has been developed for the joint and direct determination of Au and Co additives and the Sn macrocomponent [98]. The approach includes the combination of high-resolution continuum source graphite furnace atomic absorption spectrometry (HR CS GFAAS) and the application of sodium carboxylmethylcellulose (Na-CMC) for the preparation of the sample suspension. The direct determination of the analytes is possible in a suspension of powder samples and an aqueous standard solution with sr ≤ 0.04. The proposed technique may be useful for routine laboratory analyses of nanocrystalline materials. The accuracy of the results was proven by ICP MS after sample decomposition and solution analysis [98].
SnO2/MnOx nanocomposites prepared by SnO2 impregnation with Mn(acac)3 and subsequent thermal annealing [64] with various manganese contents (up to [Mn]/[Sn] = 10 mol.%) and different manganese distributions were characterized by TXRF and ICP MS. In this case, as a reagent for removing manganese from the SnO2 surface, oxalic acid was used, which formed stable complex compounds with the manganese. The obtained solutions were analyzed by the ICP MS and TXRF methods. The convergence of the results for manganese determination by these methods within an error of 5% was shown. At the same time, it was demonstrated that due to the low manganese content in the volume of the SnO2 nanocrystals, the determination of the additive concentration in samples after etching the surface was possible only using the ICP MS method in solutions obtained after sample decomposition in a microwave autoclave. A different distribution of manganese between the volume and the surface of the SnO2 crystallites was revealed depending on the total Mn concentration. The Mn (surface)/Mn (total) ratio was 0.6 ± 0.15 and 0.94 ± 0.25 for samples with a composition [Mn]/[Sn] = 1 and 10 mol.%, respectively. Thus, although the manganese was introduced as a surface modifier via the impregnation route, there was a doping effect caused by the redistribution of manganese between MnOx surface segregation and the SnO2 crystal structure. The manganese cations introduced in the SnO2 lattice acted as electroactive additives. Surface MnOx segregation had a significant influence on the sensor properties of the SnO2/MnOx nanocomposites. The presence of manganese in the total concentration of [Mn]/[Sn] = 1 mol.% increased SnO2 sensitivity towards CO and prevented signal inversion when detecting NO.

6. Conclusions

A detailed and thorough characterization of new nanocrystalline oxide materials used as a sensitive layer in semiconductor gas sensors is a necessary basis for reliably establishing the relationship between synthesis conditions, composition, structure, and sensory parameters (sensitivity and selectivity) in detecting gases of various natures. Whereas the term “composition” is complex and includes the gross quantitative elemental composition, phase composition, surface composition, electronic state of additives and main components, and distribution of additives between the volume and the surface of the metal oxide crystallites, to determine each of the abovementioned constituents, it is necessary to choose an adequate method of analysis. Besides this, the methods for determining the composition of nanocrystalline materials have a number of features. Given the small particle size and complex distribution of the additive in thin near-surface layers, the characterization of nanocomposites should be carried out using complementary methods. Along with X-ray diffraction, important information about the structure and phase composition of such systems can be obtained by scanning and transmission electron microscopy, electron diffraction, Raman spectroscopy, and Mössbauer spectroscopy. Traditional methods of surface investigation (Auger electron spectroscopy and X–ray photoelectron spectroscopy) also provide information averaged over a layer of 2–2.5 nm, which is comparable to the minimum diameter of crystal grains of 3 nm. These methods require the use of high vacuum, which creates additional difficulties in result interpretation. The approach based on the selection of a reagent for transferring an additive immobilized on the surface of tin dioxide to a solution without dissolving the SnO2 matrix, followed by the analysis of the obtained solutions by TXRF and ICP MS methods, allows one to determine the additive concentrations on the surface and in the volume of SnO2 nanocrystals with high accuracy. It should be noted that, despite the proposed approaches to the decomposition of the SnO2 matrix in a mixture of acids using an autoclave, the dissolution conditions, namely the ratio of acid concentrations and the temperature regime, must be adapted to a specific material. This time-consuming stage complicates the analysis procedure. An alternative is the use of multielement TXRF for the analysis of suspensions of undisturbed nanocomposites and the SnO2 matrix after flushing additives from the surface. The selection of an internal standard for quantitative calculations makes it possible to reproducibly determine the concentration of the additive at the level of 1 wt.%.

Author Contributions

All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Russian Science Foundation, grant number 22-13-00111, https://rscf.ru/en/project/22-13-00111/, (accessed on 22 August 2023).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ginley, D.S.; Hosono, H.; Paine, D.C. (Eds.) Handbook on Transparent Conductors; Springer: New York, NY, USA, 2011. [Google Scholar]
  2. Das, S.; Jayaraman, V. SnO2: A comprehensive review on structures and gas sensors. Prog. Mater. Sci. 2014, 66, 112–255. [Google Scholar] [CrossRef]
  3. Dalapati, G.K.; Sharma, H.; Guchhait, A.; Chakrabarty, N.; Bamola, P.; Liu, Q.; Saianand, G.; Krishna, A.M.S.; Mukhopadhyay, S.; Dey, A.; et al. Tin oxide for optoelectronic, photovoltaic and energy storage devices: A review. J. Mater. Chem. A 2021, 9, 16621–16684. [Google Scholar] [CrossRef]
  4. Hosono, H.; Mishima, Y.; Takezoe, H.; Mackenzie, K.J.D. (Eds.) Nanomaterials: From Research to Applications; Elsevier Ltd.: London, UK, 2006. [Google Scholar]
  5. Walsh, A.; Silva, L.F.D.; Wei, S.H. Origins of band gap renormalization in degenerately doped semiconductors. Phys. Rev. B 2008, 78, 075211. [Google Scholar] [CrossRef]
  6. Minami, T. Transparent conducting oxide semiconductors for transparent electrodes. Semicond. Sci. Technol. 2005, 20, S35–S44. [Google Scholar] [CrossRef]
  7. Ozel, K.; Yildiz, A. High-detectivity ultraviolet-B photodetector based on SnO2 thin film/Si heterojunction. Semicond. Sci. Technol. 2021, 36, 095001. [Google Scholar] [CrossRef]
  8. Hassun, H.K.; Hussein, B.H.; Salman, E.M.T.; Shaban, A.H. Photoelectric properties of SnO2:Ag/P–Si heterojunction photodetector. Energy Rep. 2020, 6, 46–54. [Google Scholar] [CrossRef]
  9. Yan, J.; Chen, Y.; Wang, X.; Fu, Y.; Wang, J.; Sun, J.; Dai, G.; Tao, S.; Gao, Y. High-performance solar-blind SnO2 nanowire photodetectors assembled using optical tweezers. Nanoscale 2019, 11, 2162–2169. [Google Scholar] [CrossRef]
  10. Adepu, V.; Kunchur, A.; Kolli, C.S.R.; Siddhartha, S.; Mattela, V.; Sahatiya, P. High-Performance Visible Light Photodetector Based on 1D SnO2 Nanofibers with a Ti3C2Tx (MXene) Electron Transport Layer. ACS Appl. Nano Mater. 2022, 5, 6852–6863. [Google Scholar] [CrossRef]
  11. Praveen, S.; Veeralingam, S.; Badhulika, S. A Flexible Self-Powered UV Photodetector and Optical UV Filter Based on β-Bi2O3/SnO2 Quantum Dots Schottky Heterojunction. Adv. Mater. Interfaces 2021, 8, 2100373. [Google Scholar] [CrossRef]
  12. Ye, Q.; Zhang, X.; Yao, R.; Luo, D.; Liu, X.; Zou, W.; Guo, C.; Xu, Z.; Ning, H.; Peng, J. Research and Progress of Transparent, Flexible Tin Oxide Ultraviolet Photodetector. Crystals 2021, 11, 1479. [Google Scholar] [CrossRef]
  13. Wu, P.; Wang, S.; Li, X.; Zhang, F. Advances in SnO2-based perovskite solar cells: From preparation to photovoltaic applications. J. Mater. Chem. A 2021, 9, 19554–19588. [Google Scholar] [CrossRef]
  14. Min, H.; Lee, D.Y.; Kim, J.; Kim, G.; Lee, K.S.; Paik, M.J.; Kim, Y.K.; Kim, K.S.; Kim, M.G.; Shin, T.J.; et al. Perovskite solar cells with atomically coherent interlayers on SnO2 electrodes. Nature 2021, 598, 444–450. [Google Scholar] [CrossRef] [PubMed]
  15. Chen, Y.; Meng, Q.; Zhang, L.; Han, C.; Gao, H.; Zhang, Y.; Yan, H. SnO2-based electron transporting layer materials for perovskite solar cells: A review of recent progress. J. Energy Chem. 2019, 35, 144–167. [Google Scholar] [CrossRef]
  16. Park, S.Y.; Zhu, K. Advances in SnO2 for Efficient and Stable n–i–p Perovskite Solar Cells. Adv. Mater. 2022, 34, 2110438. [Google Scholar] [CrossRef]
  17. Bondarchuk, A.N.; Corrales-Mendoza, I.; Aguilar-Martínez, J.A.; García-Pérez, U.M.; Marken, F. Porous and conductive SnO2 ceramics as a promising nanostructured substrate to host photocatalytic hematite coatings: Towards low cost solar-driven water splitting. Catal. Commun. 2023, 174, 106593. [Google Scholar] [CrossRef]
  18. Bondarchuk, A.N.; Corrales-Mendoza, I.; Aguilar-Martínez, J.A.; Tomás, S.A.; Gómez-Caiceros, D.A.; Hernández-Méndez, A.; Marken, F. A BiVO4 photoanode grown on porous and conductive SnO2 ceramics for water splitting driven by solar energy. Ceram. Int. 2020, 46, 9040–9049. [Google Scholar] [CrossRef]
  19. Mihaiu, M.S.; Scarlat, O.; Zuca, S.; Zaharescu, M. Advanced SnO2-Based Ceramics: Synthesis, Structure, Properties. In Advances in Ceramics—Synthesis and Characterization, Processing and Specific Applications; Sikalidis, C., Ed.; InTechOpen: London, UK, 2011; pp. 101–126. Available online: http://www.intechopen.com/books/advances-in-ceramics-synthesis-and-characterization-processing-andspecific-applications/advanced-sno2-based-ceramics-synthesis-structure-properties (accessed on 22 August 2023).
  20. Yan, W.; Zhang, H.; Wang, X.; You, C.; Li, Z. Characterization of electrical conductivity and temperature sensitivity of Cr/Sb-modified SnO2 ceramics. J. Mater. Sci. Mater. Electron. 2020, 31, 4040–4049. [Google Scholar] [CrossRef]
  21. Jiang, G.; Li, Z.; You, C.; Hao, W.; Ma, Z.; Zhang, H. Temperature sensitivity and electrical stability of Sb/Mn co-doped SnO2 ceramics. J. Mater. Sci. Mater. Electron. 2021, 32, 16945–16955. [Google Scholar] [CrossRef]
  22. Medvedovski, E. Tin oxide-based ceramics of high density obtained by pressureless sintering. Ceram. Int. 2017, 43, 8396–8405. [Google Scholar] [CrossRef]
  23. Masteghin, M.G.; Orlandi, M.O.; Bueno, P.R. Varistor technology based on SnO2. In Tin Oxide Materials: Synthesis, Properties, Applications; Orlandi, M.O., Ed.; Elsevier: Amsterdam, The Netherlands, 2020; pp. 321–343. [Google Scholar]
  24. Lee, J.H.; You, Y.J.; Saeed, M.A.; Kim, S.H.; Choi, S.H.; Kim, S.; Lee, S.Y.; Park, J.S.; Sim, J.W. Undoped tin dioxide transparent electrodes for efficient and cost-effective indoor organic photovoltaics. NPG Asia Mater. 2021, 13, 43. [Google Scholar] [CrossRef]
  25. Zhang, S.T.; Foldyna, M.; Roussel, H.; Consonni, V.; Pernot, E.; Schmidt-Mende, L.; Rapenne, L.; Jiménez, C.; Deschanvres, J.L.; Muñoz-Rojas, D.; et al. Tuning the properties of F:SnO2 (FTO) nanocomposites with S:TiO2 nanoparticles—Promising hazy transparent electrodes for photovoltaics applications. J. Mater. Chem. C 2017, 5, 91–102. [Google Scholar] [CrossRef]
  26. Park, S.H.; Oh, Y.K.; Lim, Y.J.; Shaozheng, C.; Lee, S.J.; Kim, H.K. Thermally stable and transparent F-doped SnO2 (FTO)/Ag/FTO films for transparent thin film heaters used in automobiles. Ceram. Int. 2023, 49, 2419–2426. [Google Scholar] [CrossRef]
  27. Seok, H.; Lee, J.; Park, J.; Lim, S.; Kim, H. Transparent Conducting Electrodes for Quantum Dots Light Emitting Diodes. Isr. J. Chem. 2019, 59, 729–746. [Google Scholar] [CrossRef]
  28. Koo, B.R.; Bae, J.W.; Ahn, H.J. Optoelectronic multifunctionality of combustion-activated fluorine-doped tin oxide films with high optical transparency. Ceram. Int. 2019, 45, 10260–10268. [Google Scholar] [CrossRef]
  29. Zum Felde, U.; Haase, M.; Weller, H. Electrochromism of Highly Doped Nanocrystalline SnO2:Sb. J. Phys. Chem. B 2000, 104, 9388–9395. [Google Scholar] [CrossRef]
  30. Goei, R.; Ong, A.J.; Tan, J.H.; Loke, J.Y.; Lua, S.K.; Mandler, D.; Magdassi, S.; Tok, A.L.Y. Nd–Nb Co-doped SnO2/α-WO3 Electrochromic Materials: Enhanced Stability and Switching Properties. ACS Omega 2021, 6, 26251–26261. [Google Scholar] [CrossRef]
  31. Tu, J.; Zhao, G.; Wang, W.; Wang, X.; Xia, X.; Gu, C. Multicolor electrochromic film based on SnO2/V2O5 core/shell structure for adaptive camouflage. J. Mater. Chem. C 2019, 7, 5702–5709. [Google Scholar]
  32. Kim, K.H.; Koo, B.R.; Ahn, H.J. Effects of Sb-doped SnO2–WO3 nanocomposite on electrochromic performance. Ceram. Int. 2019, 45, 15990–15995. [Google Scholar] [CrossRef]
  33. Jo, M.H.; Koo, B.R.; Ahn, H.J. Accelerating F-doping in transparent conducting F-doped SnO2 films for electrochromic energy storage devices. Ceram. Int. 2020, 46, 25066–25072. [Google Scholar] [CrossRef]
  34. Kong, L.B.; Ruan, S.; Xiao, Z.; Li, X.; Zhou, K.; Su, H.; Wang, C.; Zhang, T. Tin oxide-based electrochromics. In Tin Oxide Materials: Synthesis, Properties, Applications; Orlandi, M.O., Ed.; Elsevier: Amsterdam, The Netherlands, 2020; pp. 489–517. [Google Scholar]
  35. Chang, H.; Huang, C.H.; Nomura, K. Low-Temperature Solution-Processed n-Channel SnO2 Thin-Film Transistors and High-Gain Zero-VGS-Load Inverter. ACS Appl. Electron. Mater. 2021, 3, 4943–4949. [Google Scholar] [CrossRef]
  36. Cao, H.; Liang, L. Tin oxide-based thin-film transistors and their circuits. In Tin Oxide Materials: Synthesis, Properties, Applications; Orlandi, M.O., Ed.; Elsevier: Amsterdam, The Netherlands, 2020; pp. 441–476. [Google Scholar]
  37. Meng, X.; Valitova, I.; Santato, C.; Cicoira, F. Tin dioxide ion-gated transistors. In Tin Oxide Materials: Synthesis, Properties, Applications; Orlandi, M.O., Ed.; Elsevier: Amsterdam, The Netherlands, 2020; pp. 477–488. [Google Scholar]
  38. Dahl, P.I.; Barnett, A.O.; Monterrubio, F.A.; Colmenares, L.C. The use of tin oxide in fuel cells. In Tin Oxide Materials: Synthesis, Properties, Applications; Orlandi, M.O., Ed.; Elsevier: Amsterdam, The Netherlands, 2020; pp. 379–410. [Google Scholar]
  39. Notohara, H.; Urita, K.; Moriguchi, I. Tin oxide electrodes in Li and Na-ion batteries. In Tin Oxide Materials: Synthesis, Properties, Applications; Orlandi, M.O., Ed.; Elsevier: Amsterdam, The Netherlands, 2020; pp. 411–439. [Google Scholar]
  40. Manjunathan, P.; Shanbhag, G.V. Application of tin oxide-based materials in catalysis. In Tin Oxide Materials: Synthesis, Properties, Applications; Orlandi, M.O., Ed.; Elsevier: Amsterdam, The Netherlands, 2020; pp. 519–553. [Google Scholar]
  41. Staerz, A.; Suzuki, T.; Weimar, U.; Barsan, N. SnO2: The most important base material for semiconducting metal oxide-based materials. In Tin Oxide Materials: Synthesis, Properties, Applications; Orlandi, M.O., Ed.; Elsevier: Amsterdam, The Netherlands, 2020; pp. 345–377. [Google Scholar]
  42. Gaskov, A.; Rumyantseva, M.; Marikutsa, A. Tin oxide nanomaterials: Active centers and gas sensor properties. In Tin Oxide Materials: Synthesis, Properties, Applications; Orlandi, M.O., Ed.; Elsevier: Amsterdam, The Netherlands, 2020; pp. 163–218. [Google Scholar]
  43. Masuda, Y. Recent advances in SnO2 nanostructure based gas sensors. Sens. Actuators B 2022, 364, 131876. [Google Scholar] [CrossRef]
  44. Staerz, A.; Weimar, U.; Barsan, N. Current state of knowledge on metal oxide based gas sensing mechanism. Sens. Actuators B 2022, 358, 131531. [Google Scholar] [CrossRef]
  45. Rumyantseva, M.N.; Gaskov, A.M. Chemical modification of nanocrystalline metal oxides: Effect of the real structure and surface chemistry on the sensor properties. Russ. Chem. Bull. Int. Ed. 2008, 57, 1106–1125. [Google Scholar] [CrossRef]
  46. Krivetskiy, V.V.; Rumyantseva, M.N.; Gaskov, A.M. Chemical modification of nanocrystalline tin dioxide for selective gas sensors. Russ. Chem. Rev. 2013, 82, 917–941. [Google Scholar] [CrossRef]
  47. Miller, D.R.; Akbar, S.A.; Morris, P.A. Nanoscale metal oxide-based heterojunction for gas sensing: A review. Sens. Actuators B 2014, 204, 250–272. [Google Scholar] [CrossRef]
  48. Korotcenkov, G.; Cho, B.K. Metal oxide composites in conductometric gas sensors: Achievements and challenges. Sens. Actuators B 2017, 244, 182–210. [Google Scholar] [CrossRef]
  49. Zappa, D.; Galstyan, V.; Kaur, N.; Munasinghe Arachchige, H.M.M.; Sisman, O.; Comini, E. Metal oxide-based heterostructures for gas sensors—A review. Anal. Chim. Acta 2018, 1039, 1–23. [Google Scholar] [CrossRef]
  50. Walker, J.M.; Akbar, S.A.; Morris, P.A. Synergistic effects in gas sensing semiconducting oxide nanoheterostructures: A review. Sens. Actuators B 2019, 286, 624–640. [Google Scholar] [CrossRef]
  51. Marikutsa, A.; Rumyantseva, M.; Konstantinova, E.A.; Gaskov, A. The Key Role of Active Sites in the Development of Selective Metal Oxide Sensor Materials. Sensors 2021, 21, 2554. [Google Scholar] [CrossRef]
  52. Degler, D.; Weimar, U.; Barsan, N. Current Understanding of the Fundamental Mechanisms of Doped and Loaded Semiconducting Metal-Oxide-Based Gas Sensing Materials. ACS Sens. 2019, 4, 2228–2249. [Google Scholar] [CrossRef]
  53. Bordes-Richard, E. Multicomponent Oxides in Selective Oxidation of Alkanes Theoretical Acidity versus Selectivity. Top. Catal. 2008, 50, 82–89. [Google Scholar] [CrossRef]
  54. Duffy, J.A. Chemical bonding in the oxides of elements: A new appraisal. J. Solid State Chem. 1986, 62, 145–157. [Google Scholar] [CrossRef]
  55. Duffy, J.A. Ionic-covalent character of metal and nonmetal oxides. J. Phys. Chem. 2006, 110, 13245–13248. [Google Scholar] [CrossRef] [PubMed]
  56. Sanderson, R.T. An Interpretation of Bond Lengths and a Classification of Bonds. Science 1951, 114, 670–672. [Google Scholar] [CrossRef]
  57. Sanderson, R.T. The interrelationship of bond dissociation energies and contributing bond energies. J. Am. Chem. Soc. 1975, 97, 1367–1372. [Google Scholar] [CrossRef]
  58. Gleiter, H. Nanostructured materials: Basic concepts and microstructure. Acta Mater. 2000, 48, 1–29. [Google Scholar] [CrossRef]
  59. Elliot, J.A.W. Gibbsian Surface Thermodynamics. J. Phys. Chem. B 2020, 124, 10859–10878. [Google Scholar] [CrossRef]
  60. Liu, Z.-J.; Shen, Y.G. Effects of amorphous matrix on the grain growth kinetics in two-phase nanostructured films: A Monte Carlo study. Acta Mater. 2004, 52, 729–736. [Google Scholar] [CrossRef]
  61. Miodownik, M.; Holm, E.A.; Hassold, G.N. Highly parallel computer simulation of particle pinning: Zener vindicated. Scripta Mater. 2000, 42, 1173–1177. [Google Scholar] [CrossRef]
  62. Arbiol, J.; Morante, J.R.; Bouvier, P.; Pagnier, T.; Makeeva, E.; Rumyantseva, M.; Gaskov, A. SnO2/MoO3-nanostructure and alcohol detection. Sens. Actuators B 2006, 118, 156–162. [Google Scholar] [CrossRef]
  63. Vladimirova, S.A.; Rumyantseva, M.N.; Filatova, D.G.; Chizhov, A.S.; Khmelevsky, N.O.; Konstantinova, E.A.; Kozlovsky, V.F.; Marchevsky, A.V.; Karakulina, O.M.; Hadermann, J.; et al. Cobalt location in p-CoOx/n-SnO2 nanocomposites: Correlation with gas sensor performances. J. Alloys Compd. 2017, 721, 249–260. [Google Scholar] [CrossRef]
  64. Eshmakov, R.S.; Filatova, D.G.; Konstantinova, E.A.; Rumyantseva, M.N. Effect of manganese distribution on sensor properties of SnO2/MnOx nanocomposites. Nanomaterials 2023, 13, 1437. [Google Scholar] [CrossRef] [PubMed]
  65. Krivetskiy, V.; Garshev, A.; Marikutsa, A.; Ivanov, V.; Krotova, A.; Filatova, D.; Konstantinova, E.; Naberezhnyi, D.; Khmelevsky, N.; Kots, P.; et al. Enhancement of Lewis acidity of Cr-doped nanocrystalline SnO2 and its effect on surface NH3 oxidation and the sensory detection pattern. ChemPhysChem 2019, 20, 1985–1996. [Google Scholar]
  66. Rockenberger, J.; zum Felde, U.; Tischer, M.; Tröger, L.; Haase, M.; Weller, H. Near edge X-ray absorption fine structure measurements (XANES) and extended X-ray absorption fine structure measurements (EXAFS) of the valence state and coordination of antimony in doped nanocrystalline SnO2. J. Chem. Phys. 2000, 112, 4296–4304. [Google Scholar] [CrossRef]
  67. Marikutsa, A.V.; Rumyantseva, M.N.; Gaskov, A.M.; Konstantinova, E.A.; Grishina, D.A.; Deygen, D.M. CO and NH3 Sensor Properties and Paramagnetic Centers of Nanocrystalline SnO2 Modified by Pd and Ru. Thin Solid Films 2011, 520, 904–908. [Google Scholar] [CrossRef]
  68. Wolkenstein, T. Electronic Processes on Semiconductor Surfaces during Chemisorption; Springer: New York, NY, USA, 1991. [Google Scholar]
  69. Mowat, I.A.; Putyera, K.; Wang, X.; Newman, J.; Vitarelli, J. Materials Characterization and Problem Solving. Available online: http://www.ceramicindustry.com/articles/91508-materials-characterization-and-problem-solving/ (accessed on 22 August 2023).
  70. Mishra, R.K.; Zachariah, A.K.; Thomas, S. Energy-Dispersive X-Ray spectroscopy Techniques for Nanomaterials. In Microscopy Methods in Nanomaterials Characterization: Micro and Nano Technologies; Thomas, S., Tomas, R., Zachariah, A.K., Mishra, R.K., Eds.; Elsevier: Amsterdam, The Netherlands, 2017; pp. 383–405. [Google Scholar]
  71. Chen, Z.; Weyland, M.; Sang, X.; Xu, W.; Dycus, J.H.; LeBeau, J.M.; D’Alfonso, A.J.; Allen, L.J.; Findlay, S.D. Quantitative atomic resolution elemental mapping via absolute-scale energy dispersive X-ray spectroscopy. Ultramicroscopy 2016, 168, 7–16. [Google Scholar] [CrossRef]
  72. Ross, R.J. (Ed.) Microelectronics Failure Analysis Desk Reference, 6th ed.; ASM International: Detroit, MI, USA, 2011. [Google Scholar]
  73. Xu, W.; Dycus, J.H.; LeBeau, J.M. Numerical modeling of specimen geometry for quantitative energy dispersive X-ray spectroscopy. Ultramicroscopy 2017, 184A, 100–108. [Google Scholar] [CrossRef]
  74. Bilovol, V.; Herme, C.; Jacobo, S.; Cabrera, A.F. Study of magnetic behavior of Fe-doped SnO2 powders prepared by chemical method. Mater. Chem. Phys. 2012, 135, 334–339. [Google Scholar] [CrossRef]
  75. García-Tecedor, M.; Maestre, D.; Cremades, A.; Piqueras, J. Growth and characterization of Cr doped SnO2 microtubes with resonant cavity modes. J. Mater. Chem. C 2016, 4, 5709–5716. [Google Scholar] [CrossRef]
  76. Padmaja, B.; Dhanapandian, S.; Ashokkumar, K.; Krishnakumar, N. Cobalt ions doped SnO2 nanoparticles for enhanced photocatalytic and supercapacitor applications. Inorg. Chem. Commun. 2023, 155, 110948. [Google Scholar] [CrossRef]
  77. Shaikh, F.I.; Chikhale, L.P.; Mulla, I.S.; Suryavanshi, S.S. Synthesis, characterization and enhanced acetone sensing performance of Pd loaded Sm doped SnO2 nanoparticles. Ceram. Int. 2017, 43, 10307–10315. [Google Scholar] [CrossRef]
  78. Berenguer, R.; Quijada, C.; Morallón, E. Electrochemical characterization of SnO2 electrodes doped with Ru and Pt. Electrochim. Acta 2009, 54, 5230–5238. [Google Scholar] [CrossRef]
  79. Korotcenkov, G.; Gulina, L.B.; Cho, B.K.; Han, S.H.; Tolstoy, V.P. SnO2–Au nanocomposite synthesized by successive ionic layer deposition method: Characterization and application in gas sensors. Mater. Chem. Phys. 2011, 128, 433–441. [Google Scholar] [CrossRef]
  80. Liu, D.; Pan, J.; Tang, J.; Liu, W.; Bai, S.; Luo, R. Ag decorated SnO2 nanoparticles to enhance formaldehyde sensing properties. J. Phys. Chem. Solids 2019, 124, 36–43. [Google Scholar] [CrossRef]
  81. Rumyantseva, M.N.; Kovalenko, V.V.; Gaskov, A.M.; Pagnier, T.; Machon, D.; Arbiol, J.; Morante, J.R. Nanocomposites SnO2/Fe2O3: Wet chemical synthesis and nanostructure characterization. Sens. Actuators B 2005, 109, 64–74. [Google Scholar] [CrossRef]
  82. Shannon, R.D. Revised Effective Ionic Radii and Systematic Studies of Interatomic Distances in Halides and Chalcogenides. Acta Cryst. 1976, A32, 751–767. [Google Scholar] [CrossRef]
  83. Panday, M.; Upadhyay, G.K.; Purohit, L.P. Sb incorporated SnO2 nanostructured thin films for CO2 gas sensing and humidity sensing applications. J. Alloys Compd. 2022, 904, 164053. [Google Scholar] [CrossRef]
  84. Chen, X.; Liu, T.; Wu, R.; Yu, J.; Yin, X. Gas sensors based on Pd-decorated and Sb-doped SnO2 for hydrogen detection. J. Ind. Eng. Chem. 2022, 115, 491–499. [Google Scholar] [CrossRef]
  85. Kim, J.-H.; Mirzaei, A.; Kim, J.-Y.; Lee, J.-H.; Kim, H.W.; Hishita, S.; Kim, S.S. Enhancement of gas sensing by implantation of Sb-ions in SnO2 nanowires. Sens. Actuators B 2020, 304, 127307. [Google Scholar] [CrossRef]
  86. Feng, Z.; Gaiardo, A.; Valt, M.; Fabbri, B.; Casotti, D.; Krik, S.; Vanzetti, L.; Ciana, M.D.; Fioravanti, S.; Caramori, S.; et al. Investigation on Sensing Performance of Highly Doped Sb/SnO2. Sensors 2022, 22, 1233. [Google Scholar] [CrossRef]
  87. Raheem, Z.H.A.; Jebur, E.K. Effect of mixing ratio on gas sensitivity of SnO2:TiO2 thin film against oxidizing and reducing gasses. AIP Conf. Proc. 2020, 2290, 050025. [Google Scholar]
  88. Kusior, A.; Zych, L.; Zakrzewska, K.; Radecka, M. Photocatalytic activity of TiO2/SnO2 nanostructures with controlled dimensionality/complexity. Appl. Surf. Sci. 2019, 471, 973–985. [Google Scholar] [CrossRef]
  89. Łyson-Sypien, B.; Kusior, A.; Rekas, M.; Żukrowski, J.; Gajewska, M.; Michalow-Mauke, K.; Graule, T.; Radecka, M.; Zakrzewska, K. Nanocrystalline TiO2/SnO2 heterostructures for gas sensing. Beilstein J. Nanotechnol. 2017, 8, 108–122. [Google Scholar] [CrossRef] [PubMed]
  90. Krotova, A.A.; Krivetskiy, V.V.; Rumyantseva, M.N.; Filatova, D.G. Study of the Chromium Distribution in New Materials Based on Tin Dioxide by Inductively Coupled Plasma–Mass Spectrometry. Mosc. Univ. Chem. Bull. 2019, 74, 10–13. [Google Scholar] [CrossRef]
  91. Huang, J.Q.; Liu, Y.K.; Wu, Y.M.; Li, X.M. Influence of Mn Doping on the Sensing Properties of SnO2 Nanobelt to Ethanol. Am. J. Anal. Chem. 2017, 8, 60–71. [Google Scholar] [CrossRef]
  92. Gandhi, T.I.; Babu, R.R.; Ramamurthi, K.; Arivanandhan, M. Effect of Mn doping on the electrical and optical properties of SnO2 thin films deposited by chemical spray pyrolysis technique. Thin Solid Films 2016, 598, 195–203. [Google Scholar] [CrossRef]
  93. Sun, Q.; Xu, X.; Peng, H.; Fang, X.; Liu, W.; Ying, J.; Yu, F.; Wang, X. SnO2-based solid solutions for CH4 deep oxidation: Quantifying the lattice capacity of SnO2 using an X-ray diffraction extrapolation method. Chin. J. Catal. 2016, 37, 1293–1302. [Google Scholar] [CrossRef]
  94. Kovalenko, V.V.; Rumyantseva, M.N.; Fabritchnyi, P.B.; Gaskov, A.M. The unusual distribution of the constituants in the (Fe2O3)0.8(SnO2)0.2 nanocomposite evidenced by 57Fe and 119Sn Mössbauer spectroscopy. Mendeleev Commun. 2004, 14, 140–141. [Google Scholar] [CrossRef]
  95. Han, Z.; Tang, Y.; Lu, G.; Qi, Y.; Wu, H.; Yang, Z.; Han, H.; Zhang, X.; Wu, L.; Wang, Z.; et al. Transition metal elements-doped SnO2 for ultrasensitive and rapid ppb-level formaldehyde sensing. Heliyon 2023, 9, e13486. [Google Scholar] [CrossRef]
  96. Soussi, L.; Garmim, T.; Karzazi, O.; Rmili, A.; El Bachiri, A.; Louardi, A.; Erguig, H. Effect of (Co, Fe, Ni) doping on structural, optical and electrical properties of sprayed SnO2 thin film. Surf. Interfaces 2020, 19, 100467. [Google Scholar] [CrossRef]
  97. Zhang, H.; Zhang, H.; Man, J.; Chen, C. Preparation of high-performance Fe-doped SnO2 humidity sensor and its application in respiration detection. Sens. Actuators A 2023, 362, 114644. [Google Scholar] [CrossRef]
  98. Filatova, D.G.; Eskina, V.V.; Baranovskaya, V.B.; Vladimirova, S.A.; Gaskov, A.M.; Rumyantseva, M.N.; Karpov, Y.A. Determination of gold and cobalt dopants in advanced materials based on tin oxide by slurry sampling high-resolution continuum source graphite furnace atomic absorption spectrometry. Spectrochim. Acta B 2018, 140, 1–4. [Google Scholar] [CrossRef]
  99. Zhou, S.; Wang, H.; Hu, J.; Lv, T.; Rong, Q.; Zhang, Y.; Zi, B.; Chen, M.; Zhang, D.; Wei, J.; et al. Formaldehyde gas sensor with extremely high response employing cobalt-doped SnO2 ultrafine nanoparticles. Nanoscale Adv. 2022, 4, 824–836. [Google Scholar] [CrossRef]
  100. Korotcenkov, G.; Boris, I.; Brinzari, V.; Han, S.H.; Cho, B.K. The role of doping effect on the response of SnO2-based thin film gas sensors: Analysis based on the results obtained for Co-doped SnO2 films deposited by spray pyrolysis. Sens. Actuators B 2013, 182, 112–124. [Google Scholar] [CrossRef]
  101. Chen, K.; Zhou, Y.; Jin, R.; Wang, T.; Liu, F.; Wang, C.; Yan, X.; Sun, P.; Lu, G. Gas sensor based on cobalt-doped 3D inverse opal SnO2 for air quality monitoring. Sens. Actuators B 2022, 350, 130807. [Google Scholar] [CrossRef]
  102. Li, J.; Zheng, M.; Yang, M.; Zhang, X.; Cheng, X.; Zhou, X.; Gao, S.; Xu, Y.; Huo, L. Three-in-one Ni doped porous SnO2 nanorods sensor: Controllable oxygen vacancies content, surface site activation and low power consumption for highly selective NO2 monitoring. Sens. Actuators B 2023, 382, 133550. [Google Scholar] [CrossRef]
  103. Gu, C.; Guan, W.; Liu, X.; Gao, L.; Wang, L.; Shim, J.-J.; Huang, J. Controlled synthesis of porous Ni-doped SnO2 microstructures and their enhanced gas sensing properties. J. Alloys Compd. 2017, 692, 855–864. [Google Scholar] [CrossRef]
  104. Inderan, V.; Arafat, M.M.; Kumar, S.; Haseeb, A.S.M.A.; Jiang, Z.-T.; Altarawneh, M.; Lee, H.L. Study of structural properties and defects of Ni-doped SnO2 nanorods as ethanol gas sensors. Nanotechnology 2017, 28, 265702. [Google Scholar] [CrossRef]
  105. Manikandan, V.; Petrila, I.; Vigneselvan, S.; Mane, R.S.; Vasile, B.; Dharmavarapu, R.; Lundgaard, S.; Juodkazis, S.; Chandrasekarang, J. A reliable chemiresistive sensor of nickel-doped tin oxide (Ni-SnO2) for sensing carbon dioxide gas and humidity. RSC Adv. 2020, 10, 3796–3804. [Google Scholar] [CrossRef]
  106. Zhang, L.; He, J.; Jiao, W. Synthesis and gas sensing performance of NiO decorated SnO2 vertical standing nanotubes composite thin films. Sens. Actuators B 2019, 281, 326–334. [Google Scholar] [CrossRef]
  107. Bhardwaj, N.; Pandey, A.; Satpati, B.; Tomar, M.; Gupta, V.; Mohapatra, S. Enhanced CO gas sensing properties of Cu doped SnO2 nanostructures prepared by a facile wet chemical method. Phys. Chem. Chem. Phys. 2016, 18, 18846. [Google Scholar] [CrossRef] [PubMed]
  108. Chen, T.; Sun, J.; Xue, N.; Wang, W.; Luo, Z.; Liang, Q.; Zhou, T.; Quan, H.; Cai, H.; Tang, K.; et al. Cu-doped SnO2/rGO nanocomposites for ultrasensitive H2S detection at low temperature. Microsyst. Nanoeng. 2023, 9, 69. [Google Scholar] [CrossRef] [PubMed]
  109. Xu, X.; Tong, Y.; Zhang, J.; Fang, X.; Xu, J.; Liu, F.; Liu, J.; Zhong, W.; Lebedeva, O.E.; Wang, X. Investigation of lattice capacity effect on Cu2+-doped SnO2 solid solution catalysts to promote reaction performance toward NOx-SCR with NH3. Chin. J. Catal. 2020, 41, 877–888. [Google Scholar] [CrossRef]
  110. Wang, W.; Liu, Y.; Liu, S. SnO2 Nanostructure with Well-Engineered Crystal Facets by Zn Doping for Chemical Sensing Applications. Cryst. Growth Des. 2020, 20, 2742–2752. [Google Scholar] [CrossRef]
  111. Minau, S.; Atkinson, I.; Mocioiu, O.; Toader, A.; Tenea, E.; Zaharescu, M. Phase formation mechanism in the ZnO-SnO2 binary system. Rev. Roum. Chim. 2011, 56, 465–472. [Google Scholar]
  112. Boomashri, M.; Perumal, P.; Khan, A.; El-Toni, A.M.; Ansari, A.A.; Gupta, R.K.; Murahari, P.; Kumar, K.D.A. Zinc influence on nanostructured tin oxide (SnO2) films as ammonia sensor at room temperature. Surf. Interfaces 2021, 25, 101195. [Google Scholar] [CrossRef]
  113. Khamfoo, K.; Wisitsoraat, A.; Punginsang, M.; Tuantranont, A.; Liewhiran, C. Selectivity towards acetylene gas of flame-spray-made Nb-substituted SnO2 particulate thick films. Sens. Actuators B 2021, 349, 130808. [Google Scholar] [CrossRef]
  114. Gao, S.; Guan, H.; Wang, H.; Yang, X.; Yang, W.; Li, Q. Creation of SnxNb1−xO2 solid solution through heavy Nb-doping in SnO2 to boost its photocatalytic CO2 reduction to C2+ products under simulated solar illumination. J. Adv. Ceram. 2022, 11, 1404–1416. [Google Scholar] [CrossRef]
  115. Zhu, Y.; Shan, W.; Lian, Z.; Liu, J.; Zhang, Y.; He, H. Effects of impregnation sequence on the NH3-SCR activity and hydrothermal stability of a Ce-Nb/SnO2 catalyst. J. Environ. Sci. 2023, 138, 450–457. [Google Scholar] [CrossRef]
  116. Marikutsa, A.V.; Rumyantseva, M.N.; Konstantinova, E.A.; Shatalova, T.B.; Gaskov, A.M. Active sites on nanocrystalline tin dioxide surface: Effect of palladium and ruthenium oxides clusters. J. Phys. Chem. C 2014, 118, 21541–21549. [Google Scholar] [CrossRef]
  117. Kou, X.; Meng, F.; Chen, K.; Wang, T.; Sun, P.; Liu, F.; Yan, X.; Sun, Y.; Liu, F.; Shimanoe, K.; et al. High performance acetone gas sensor based on Ru-doped SnO2 nanofibers. Sens. Actuators B 2020, 320, 128292. [Google Scholar] [CrossRef]
  118. Li, J.; Xian, J.; Wang, W.; Cheng, K.; Zeng, M.; Zhang, A.; Wu, S.; Gao, X.; Lu, X.; Liu, J.-M. Ultrafast response and high-sensitivity acetone gas sensor based on porous hollow Ru-doped SnO2 nanotubes. Sens. Actuators B 2022, 352, 131061. [Google Scholar] [CrossRef]
  119. Senthilkumar, P.; Raja, S.; Babu, R.R.; Vasuki, G. Influence of Ru doping on the structural, morphological, optical, electrical and optoelectronic properties of SnO2 thin films. J. Phys. Chem. Solids 2023, 174, 111117. [Google Scholar] [CrossRef]
  120. Xu, X.; Liu, F.; Huang, J.; Luo, W.; Yu, J.; Fang, X.; Lebedeva, O.E.; Wang, X. The influence of RuO2 distribution and dispersion on the reactivity of RuO2-SnO2 composite oxide catalysts probed by CO oxidation. ChemCatChem 2019, 11, 2473–2483. [Google Scholar] [CrossRef]
  121. Batzill, M.; Diebold, U. The surface and materials science of tin oxide. Prog. Surf. Sci. 2005, 79, 47–154. [Google Scholar] [CrossRef]
  122. Koziej, D.; Hubner, M.; Barsan, N.; Weimar, U.; Sikorazc, M.; Grunwaldt, J.-D. Operando X-ray absorption spectroscopy studies on Pd-SnO2 based sensors. Phys. Chem. Chem. Phys. 2009, 11, 8620–8625. [Google Scholar] [CrossRef] [PubMed]
  123. Marikutsa, A.V.; Rumyantseva, M.N.; Yashina, L.V.; Gaskov, A.M. Role of surface hydroxyl groups in promoting room temperature CO sensing by Pd-modified nanocrystalline SnO2. J. Solid State Chem. 2010, 183, 2389–2399. [Google Scholar] [CrossRef]
  124. Gschwend, P.M.; Schenk, F.M.; Gogos, A.; Pratsinis, S.E. Acetone Sensing and Catalytic Conversion by Pd-Loaded SnO2. Materials 2021, 14, 5921. [Google Scholar] [CrossRef] [PubMed]
  125. Schierbaum, K.D.; Kirner, U.K.; Geiger, J.F.; Göpel, W. Schottky-barrier and conductivity gas sensors based upon Pd/SnO2 and Pt/TiO2. Sens. Actuators B 1991, 4, 87–94. [Google Scholar] [CrossRef]
  126. Rettenmaier, C.; Arán-Ais, R.M.; Timoshenko, J.; Rizo, R.; Jeon, H.S.; Kühl, S.; Chee, S.W.; Bergmann, A.; Cuenya, B.R. Enhanced Formic Acid Oxidation over SnO2-decorated Pd Nanocubes. ACS Catal. 2020, 10, 14540–14551. [Google Scholar] [CrossRef]
  127. Gaidi, M.; Hazemann, J.L.; Matko, I.; Chenevier, B.; Rumyantseva, M.; Gaskov, A.; Labeau, M. Role of Pt aggregates in Pt/SnO2 thin films used as gas sensors: Investigation of the catalytic effect. J. Electrochem. Soc. 2000, 147, 3131–3139. [Google Scholar] [CrossRef]
  128. Kutukov, P.; Rumyantseva, M.; Krivetskiy, V.; Filatova, D.; Batuk, M.; Hadermann, J.; Khmelevsky, N.; Aksenenko, A.; Gaskov, A. Influence of mono- and bimetallic PtOx, PdOx, PtPdOx clusters on CO sensing by SnO2 based gas sensors. Nanomaterials 2018, 8, 917. [Google Scholar] [CrossRef]
  129. Tammanoon, N.; Wisitsoraat, A.; Phokharatkul, D.; Tuantranont, A.; Phanichphant, S.; Yordsri, V.; Liewhiran, C. Highly sensitive acetone sensors based on flame-spray-made La2O3-doped SnO2 nanoparticulate thick films. Sens. Actuators B 2018, 262, 245–262. [Google Scholar] [CrossRef]
  130. Choi, J.Y.; Oh, T.S. CO sensitivity of La2O3-doped SnO2 thick film gas sensor. Thin Solid Films 2013, 547, 230–234. [Google Scholar] [CrossRef]
  131. Leangtanom, P.; Wisitsoraat, A.; Jaruwongrungsee, K.; Chanlek, N.; Phanichphant, S.; Kruefu, V. Highly sensitive and selective ethylene gas sensors based on CeOx-SnO2 nanocomposites prepared by a Co-precipitation method. Mater. Chem. Phys. 2020, 254, 123540. [Google Scholar] [CrossRef]
  132. Wang, Y.; Li, H.; Huang, D.; Wang, X.; Cai, L.; Chen, Y.; Wang, W.; Song, Y.; Han, G.; Zhen, B. A high-performance ethanol gas sensor based on Ce-doped SnO2 nanomaterials prepared by the Pechini method. Mater. Sci. Semicond. Proc. 2022, 137, 106188. [Google Scholar] [CrossRef]
  133. Kotchasak, N.; Wisitsoraat, A.; Tuantranont, A.; Phanichphant, S.; Yordsri, V.; Liewhiran, C. Highly sensitive and selective detection of ethanol vapor using flame-spray-made CeOx-doped SnO2 nanoparticulate thick films. Sens. Actuators B 2018, 255, 8–21. [Google Scholar] [CrossRef]
  134. Pagnier, T.; Boulova, M.; Galerie, A.; Gaskov, A.; Lucazeau, G. Reactivity of SnO2 –CuO nanocrystalline materials with H2S: A coupled electrical and Raman spectroscopic study. Sens. Actuators B 2000, 71, 134–139. [Google Scholar] [CrossRef]
  135. Nasriddinov, A.; Platonov, V.; Garshev, A.; Rumyantseva, M. Low temperature HCHO detection by SnO2/TiO2@Au and SnO2/TiO2@Pt: Understanding by in-situ DRIFT spectroscopy. Nanomaterials 2021, 11, 2049. [Google Scholar] [CrossRef]
  136. Chen, G.; Ji, S.; Li, H.; Kang, X.; Chang, S.; Wang, Y.; Yu, G.; Lu, J.; Claverie, J.; Sang, Y.; et al. High-Energy Faceted SnO2-Coated TiO2 Nanobelt Heterostructure for Near-Ambient Temperature-Responsive Ethanol Sensor. ACS Appl. Mater. Interfaces 2015, 7, 24950–24956. [Google Scholar] [CrossRef]
  137. Ng, S.; Prášek, J.; Zazpe, R.; Pytlíček, Z.; Spotz, Z.; Rodriguez Pereira, J.; Michalička, J.; Přikryl, J.; Krbal, M.; Sopha, H.; et al. Atomic Layer Deposition of SnO2-Coated Anodic One-Dimensional TiO2 Nanotube Layers for Low Concentration NO2 Sensing. ACS Appl. Mater. Interfaces 2020, 12, 33386–33396. [Google Scholar] [CrossRef] [PubMed]
  138. Marzec, A.; Radecka, M.; Maziarz, W.; Kusior, A.; Pędzich, Z. Structural, optical and electrical properties of nanocrystalline TiO2, SnO2 and their composites obtained by the sol–gel method. J. Eur. Ceram. Soc. 2016, 36, 2981–2989. [Google Scholar] [CrossRef]
  139. Xu, H.; Ju, J.; Li, W.; Zhang, J.; Wang, J.; Cao, B. Superior triethylamine-sensing properties based on TiO2/SnO2 n–n heterojunction nanosheets directly grown on ceramic tubes. Sens. Actuators B 2016, 228, 634–642. [Google Scholar] [CrossRef]
  140. Li, F.; Gao, X.; Wang, R.; Zhang, T.; Lu, G. Study on TiO2-SnO2 core-shell heterostructure nanofibers with different work function and its application in gas sensor. Sens. Actuators B 2017, 248, 812–819. [Google Scholar] [CrossRef]
  141. Nowak, P.; Maziarz, W.; Rydosz, A.; Kowalski, K.; Ziąbka, M.; Zakrzewska, K. SnO2/TiO2 Thin Film n-n Heterostructures of Improved Sensitivity to NO2. Sensors 2020, 20, 6830. [Google Scholar] [CrossRef] [PubMed]
  142. Ko, W.C.; Kim, K.M.; Kwon, Y.J.; Choi, H.; Park, J.K.; Jeong, Y.K. ALD-assisted synthesis of V2O5 nanoislands on SnO2 nanowires for improving NO2 sensing performance. Appl. Surf. Sci. 2020, 509, 144821. [Google Scholar] [CrossRef]
  143. Staerz, A.; Gao, X.; Cetmi, F.; Ming, Z.; Weimar, U.; Zhang, T.; Barsan, N. Dominant Role of Heterojunctions in Gas Sensing with Composite Materials. ACS Appl. Mater. Interfaces 2020, 12, 21127–21132. [Google Scholar] [CrossRef]
  144. Jeong, S.-Y.; Moon, Y.K.; Kim, T.-H.; Park, S.-W.; Kim, K.B.; Kang, Y.C.; Lee, J.-H. A New Strategy for Detecting Plant Hormone Ethylene Using Oxide Semiconductor Chemiresistors: Exceptional Gas Selectivity and Response Tailored by Nanoscale Cr2O3 Catalytic Overlayer. Adv. Sci. 2020, 7, 1903093. [Google Scholar] [CrossRef]
  145. An, Y.; Wang, T.; Li, T.; Yang, H.; Yu, H.; Xia, L.; Huang, X. Ultra-high gas sensing properties of porous Cr2O3@SnO2 composite toward NOx based on the p-p heterostructure. Ceram. Int. 2022, 48, 21982–21987. [Google Scholar] [CrossRef]
  146. Gao, X.; Zhou, Q.; Lu, Z.; Xu, L.; Zhang, Q.; Zeng, W. Synthesis of Cr2O3 Nanoparticle-Coated SnO2 Nanofibers and C2H2 Sensing Properties. Front. Mater. 2019, 6, 63. [Google Scholar] [CrossRef]
  147. Singh, A.; Verma, A.; Yadav, B.C. MnO2-SnO2 Based Liquefied Petroleum Gas Sensing Device for Lowest Explosion Limit Gas Concentration. ECS Sens. Plus 2022, 1, 025201. [Google Scholar] [CrossRef]
  148. Bigiani, L.; Zappa, D.; Maccato, C.; Gasparotto, A.; Sada, C.; Comini, E.; Barreca, D. Hydrogen Gas Sensing Performances of p-Type Mn3O4 Nanosystems: The Role of Built-in Mn3O4/Ag and Mn3O4/SnO2 Junctions. Nanomaterials 2020, 10, 511. [Google Scholar] [CrossRef] [PubMed]
  149. Yin, X.-T.; Wu, S.-S.; Dastan, D.; Nie, N.; Liu, Y.; Li, Z.-G.; Zhou, Y.-W.; Li, J.; Faik, A.; Shan, K.; et al. Sensing selectivity of SnO2-Mn3O4 nanocomposite sensors for the detection of H2 and CO gases. Surf. Interfaces 2021, 25, 101190. [Google Scholar] [CrossRef]
  150. Jayababu, N.; Poloju, M.; Ramana Reddy, M.V. Facile synthesis of SnO2-Fe2O3 core-shell nanostructures and their 2-methoxyethanol gas sensing characteristics. J. Alloys Compd. 2019, 780, 523–533. [Google Scholar] [CrossRef]
  151. Sun, P.; Zhou, X.; Wang, C.; Shimanoe, K.; Lu, G.; Yamazoe, N. Hollow SnO2/α-Fe2O3 spheres with a double-shell structure for gas sensors. J. Mater. Chem. A 2014, 2, 1302–1308. [Google Scholar] [CrossRef]
  152. Salah, B.; Ayesh, A.I. Fabrication of H2S sensitive gas sensors formed of SnO2–Fe2O3 composite nanoparticles. Mater. Chem. Phys. 2021, 266, 124597. [Google Scholar] [CrossRef]
  153. Zhang, B.; Fu, W.; Meng, X.; Ruan, A.; Su, P.; Yang, H. Enhanced ethanol sensing properties based on spherical-coral-like SnO2 nanorods decorated with α-Fe2O3 nanocrystallites. Sens. Actuators B 2018, 261, 505–514. [Google Scholar] [CrossRef]
  154. Yan, S.; Xue, J.; Wu, Q. Synchronous synthesis and sensing performance of α-Fe2O3/SnO2 nanofiber heterostructures for conductometric C2H5OH detection. Sens. Actuators B 2018, 275, 322–331. [Google Scholar] [CrossRef]
  155. Li, X.-B.; Wang, D.; Gao, C.; Zhao, X.; Sun, S.; Ren, Y.; Zhang, Q.; Zhang, J.; Dang, W.; Zhao, Y.-X.; et al. Isobutanol gas sensor based on Fe2O3–SnO2 heterostructure nanostructure. Vacuum 2023, 218, 112624. [Google Scholar] [CrossRef]
  156. Kwon, Y.J.; Na, H.G.; Kang, S.Y.; Choi, M.S.; Bang, J.H.; Kim, T.W.; Mirzaei, A.; Kim, H.W. Attachment of Co3O4 layer to SnO2 nanowires for enhanced gas sensing properties. Sens. Actuators B 2017, 239, 180–182. [Google Scholar] [CrossRef]
  157. Kim, H.; Cai, Z.; Chang, S.-P.; Park, S. Improved sub-ppm acetone sensing properties of SnO2 nanowire-based sensor by attachment of Co3O4 nanoparticles. J. Mater. Res. Technol. 2020, 9, 1129–1136. [Google Scholar] [CrossRef]
  158. Raza, M.H.; Kaur, N.; Comini, E.; Pinna, N. Toward Optimized Radial Modulation of the Space-Charge Region in One-Dimensional SnO2−NiO Core−Shell Nanowires for Hydrogen Sensing. ACS Appl. Mater. Interfaces 2020, 12, 4594–4606. [Google Scholar] [CrossRef]
  159. Niu, G.; Zhao, C.; Gong, G.; Yang, Z.; Leng, X.; Wang, F. NiO nanoparticle-decorated SnO2 nanosheets for ethanol sensing with enhanced moisture resistance. Microsyst. Nanoeng. 2019, 5, 21. [Google Scholar] [CrossRef] [PubMed]
  160. Jayababu, N.; Poloju, M.; Shruthi, J.; Ramana Reddy, M.V. Semi shield driven p-n heterostructures and their role in enhancing the room temperature ethanol gas sensing performance of NiO/SnO2 nanocomposites. Ceram. Int. 2019, 45, 15134–15142. [Google Scholar] [CrossRef]
  161. Ju, D.; Xu, H.; Xu, Q.; Gong, H.; Qiu, Z.; Guo, J.; Zhang, J.; Cao, B. High triethylamine-sensing properties of NiO/SnO2 hollow sphere P–N heterojunction sensors. Sens. Actuators B 2015, 215, 39–44. [Google Scholar] [CrossRef]
  162. Park, K.-R.; Cho, H.-B.-; Lee, J.; Song, Y.; Kim, W.-B.; Cho, Y.-H. Design of highly porous SnO2-CuO nanotubes for enhancing H2S gas sensor performance. Sens. Actuators B 2020, 302, 127179. [Google Scholar] [CrossRef]
  163. Kumar, A.; Kumar Shringi, A.; Kumar, M. RF sputtered CuO anchored SnO2 for H2S gas sensor. Sens. Actuators B 2022, 370, 132417. [Google Scholar] [CrossRef]
  164. Duoc, V.T.; Hung, C.M.; Nguyen, H.; Duy, N.V.; Hieu, N.V.; Hoa, N.D. Room temperature highly toxic NO2 gas sensors based on rootstock/scion nanowires of SnO2/ZnO, ZnO/SnO2, SnO2/SnO2 and, ZnO/ZnO. Sens. Actuators B 2021, 348, 130652. [Google Scholar] [CrossRef]
  165. Li, H.; Chu, S.; Ma, Q.; Li, H.; Che, Q.; Wang, J.; Wang, G.; Yang, P. Multilevel Effective Heterojunctions Based on SnO2/ZnO 1D Fibrous Hierarchical Structure with Unique Interface Electronic Effects. ACS Appl. Mater. Interfaces 2019, 11, 31551–31561. [Google Scholar] [CrossRef]
  166. Zheng, X.; Fan, H.; Wang, H.; Yan, B.; Ma, J.; Wang, W.; Yadav, A.K.; Dong, W.; Wang, S. ZnO–SnO2 nano-heterostructures with high-energy facets for high selective and sensitive chlorine gas sensor. Ceram. Int. 2020, 46, 27499–27507. [Google Scholar] [CrossRef]
  167. Mao, L.-W.; Zhu, L.-Y.; Wu, T.T.; Xu, L.; Jin, X.-H.; Lu, H.-L. Excellent long-term stable H2S gas sensor based on Nb2O5/SnO2 core-shell heterostructure nanorods. Appl. Surf. Sci. 2022, 602, 154339. [Google Scholar] [CrossRef]
  168. Li, C.; Kim, K.; Fuchigami, T.; Asaka, T.; Kakimoto, K.; Masuda, Y. Acetone gas sensor based on Nb2O5@SnO2 hybrid structure with high selectivity and ppt-level sensitivity. Sens. Actuators B 2023, 393, 134144. [Google Scholar] [CrossRef]
  169. Yu, J.; Zhao, D.; Xu, X.; Wang, X.; Zhang, N. Study on RuO2/SnO2: Novel and Active Catalysts for CO and CH4 Oxidation. ChemCatChem 2012, 4, 1122–1132. [Google Scholar] [CrossRef]
  170. Chung, F.H. Quantitative interpretation of X-ray diffraction patterns of mixtures. II. Adiabatic principle of X-ray diffraction analysis of mixtures. J. Appl. Cryst. 1974, 7, 526–530. [Google Scholar] [CrossRef]
  171. Hu, K.; Li, Y.; Ge, C.; Bai, L.; Liu, G.; Qiao, G.; Kang, S.G.; Kim, E.J.; Wang, M. Room-temperature ppb-level NO2 sensitivity of three-dimensional ordered macroporous Au-loaded SnO2 under intermittent UV light irradiation. Sens. Actuators B 2023, 387, 133786. [Google Scholar] [CrossRef]
  172. Ischenko, A.A.; Lazov, M.A.; Mironova, E.V.; Putin, A.Y.; Ionov, A.M.; Storozhenko, P.A. Analysis of nanoparticles and nanomaterials using X-ray photoelectron spectroscopy. Fine Chem. Technol. 2023, 18, 135–167. [Google Scholar] [CrossRef]
  173. Krishna, D.N.G.; Philip, J. Review on surface characterization applications of X-ray photoelectron spectroscopy (XPS): Recent developments and challenges. Appl. Surf. Sci. Adv. 2022, 12, 100332. [Google Scholar] [CrossRef]
  174. Barsan, N.; Weimar, U. Conduction model of metal oxide gas sensors. J. Electroceram. 2001, 7, 143–167. [Google Scholar] [CrossRef]
  175. Gunji, S.; Jukei, M.; Shimotsuma, Y.; Miura, K.; Suematsu, K.; Watanabe, K.; Shimanoe, K. Unexpected gas sensing property of SiO2/SnO2 core-shell nanofibers in dry and humid conditions. J. Mater. Chem. C 2017, 5, 6369–6376. [Google Scholar] [CrossRef]
  176. Gulevich, D.; Rumyantseva, M.; Marikutsa, A.; Shatalova, T.; Konstantinova, E.; Gerasimov, E.; Gaskov, A. Nanocomposites SnO2/SiO2: SiO2 Impact on the Active Centers and Conductivity Mechanism. Materials 2019, 12, 3618. [Google Scholar] [CrossRef]
  177. Gulevich, D.; Rumyantseva, M.; Gerasimov, E.; Marikutsa, A.; Krivetskiy, V.; Shatalova, T.; Khmelevsky, N.; Gaskov, A. Nanocomposites SnO2/SiO2 for CO gas sensors: Microstructure and reactivity in the interaction with the gas phase. Materials 2019, 12, 1096. [Google Scholar] [CrossRef] [PubMed]
  178. Marshall, A.T.; Haverkamp, R.G. Electrocatalytic activity of IrO2–RuO2 supported on Sb-doped SnO2 nanoparticles. Electrochim. Acta 2010, 55, 1978–1984. [Google Scholar] [CrossRef]
  179. Costa, I.M.; Colmenares, Y.N.; Pizani, P.S.; Leite, E.R.; Chiquito, A.J. Sb doping of VLS synthesized SnO2 nanowires probed by Raman and XPS spectroscopy. Chem. Phys. Lett. 2018, 695, 125–130. [Google Scholar] [CrossRef]
  180. Wang, S.; Zhao, Y.; Huang, J.; Wang, Y.; Ren, H.; Wu, S.; Zhang, S.; Huang, W. Low-temperature CO gas sensors based on Au/SnO2 thick film. Appl. Surf. Sci. 2007, 253, 3057–3061. [Google Scholar] [CrossRef]
  181. Neri, G.; Bonavita, A.; Milone, C.; Galvagno, S. Role of the Au oxidation state in the CO sensing mechanism of Au/iron oxide-based gas sensors. Sens. Actuators B 2003, 93, 402. [Google Scholar] [CrossRef]
  182. Li, J.; Wang, L.; Cheng, X.; Luo, D.; Huang, B.; Sun, S.; Li, X.; Yang, Z. Low-temperature and high-sensitivity Au-decorated thin-walled SnO2 nanotubes sensor for ethanol detection. Mater. Today Commun. 2023, 37, 107217. [Google Scholar] [CrossRef]
  183. Gulevich, D.; Rumyantseva, M.; Gerasimov, E.; Khmelevsky, N.; Tsvetkova, E.; Gaskov, A. Synergy effect of Au and SiO2 modification on SnO2 sensor properties in VOCs detection in humid air. Nanomaterials 2020, 10, 813. [Google Scholar] [CrossRef]
  184. Fujitani, T.; Nakamura, I.; Haruta, M. Role of Water in CO Oxidation on Gold Catalysts. Catal. Lett. 2014, 144, 1475–1486. [Google Scholar] [CrossRef]
  185. Biesinger, M.C.; Payne, B.P.; Grosvenor, A.P.; Lau, L.W.M.; Gerson, A.R.; Smart, R.S.C. Resolving surface chemical states in XPS analysis of first row transition metals, oxides and hydroxides: Cr, Mn, Fe, Co and Ni. Appl. Surf. Sci. 2011, 257, 2717–2730. [Google Scholar] [CrossRef]
  186. Wang, T.; Cheng, L. Hollow hierarchical TiO2-SnO2-TiO2 composite nanofibers with increased active-sites and charge transfer for enhanced acetone sensing performance. Sens. Actuators B 2021, 334, 129644. [Google Scholar] [CrossRef]
  187. Stranick, M.A. Mn2O3 by XPS. Surf. Sci. Spectra 1999, 6, 39–46. [Google Scholar] [CrossRef]
  188. Stranick, M.A. MnO2 by XPS. Surf. Sci. Spectra 1999, 6, 31–38. [Google Scholar] [CrossRef]
  189. Tan, Y.; Zhang, J. Highly sensitive ethanol gas sensors based on Co-doped SnO2 nanobelts and pure SnO2 nanobelts. Physica E 2023, 147, 115604. [Google Scholar] [CrossRef]
  190. Frolov, D.D.; Kotovshchikov, Y.N.; Morozov, I.V.; Boltalin, A.I.; Fedorova, A.A.; Marikutsa, A.V.; Rumyantseva, M.N.; Gaskov, A.M.; Sadovskaya, E.M.; Abakumov, A.M. Oxygen exchange on nanocrystalline tin dioxide modified by palladium. J. Solid State Chem. 2012, 186, 1–8. [Google Scholar] [CrossRef]
  191. Bahrami, B.; Khodadadi, A.; Kazemeini, M.; Mortazavi, Y. Enhanced CO sensitivity and selectivity of gold nanoparticles-doped SnO2 sensor in presence of propane and methane. Sens. Actuators B 2008, 133, 352–356. [Google Scholar] [CrossRef]
  192. Cerdà Belmonte, J.; Manzano, J.; Arbiol, J.; Cirera, A.; Puigcorbe, J.; Vilà, A.; Sabaté, N.; Gràcia, I.; Cané, C.; Morante, J.R. Micromachined twin gas sensor for CO and O2 quantification based on catalytically modified nano-SnO2. Sens. Actuators B 2006, 114, 881–892. [Google Scholar] [CrossRef]
  193. Marikutsa, A.; Krivetskiy, V.; Yashina, L.; Rumyantseva, M.; Konstantinova, E.; Ponzoni, A.; Comini, E.; Abakumov, A.; Gaskov, A. Catalytic impact of RuOx clusters to high ammonia sensitivity of tin dioxide. Sens. Actuators B 2012, 175, 186–193. [Google Scholar] [CrossRef]
  194. Safonova, O.V.; Delabouglise, G.; Chenevier, B.; Gaskov, A.M.; Labeau, M. CO and NO2 gas sensitivity of nanocrystalline tin dioxide thin films doped with Pd, Ru and Rh. Mater. Sci. Eng. C 2002, 21, 105–111. [Google Scholar] [CrossRef]
  195. Liu, T.J.; Jin, Z.G.; Feng, L.R.; Wang, T. Conducting antimony-doped tin oxide films derived from stannous oxalate by aqueous sol–gel method. Appl. Surf. Sci. 2008, 254, 6547–6553. [Google Scholar] [CrossRef]
  196. Noonuruk, R.; Vittayakorn, N.; Mekprasart, W.; Sritharathikhun, J.; Pecharapa, W. Sb-Doped SnO2 Nanoparticles Synthesized by Sonochemical-Assisted Precipitation Process. J. Nanosci. Nanotechnol. 2015, 15, 2564–2569. [Google Scholar] [CrossRef]
  197. Liu, J.; Liu, X.; Zhai, Z.; Jin, G.; Jiang, Q.; Zhao, Y.; Luo, C.; Quan, L. Evaluation of depletion layer width and gas-sensing properties of antimony-doped tin oxide thin film sensors. Sens. Actuators B 2015, 220, 1354–1360. [Google Scholar] [CrossRef]
  198. Yao, P. Effects of Sb doping level on the properties of Ti/SnO2-Sb electrodes prepared using ultrasonic spray pyrolysis. Desalination 2011, 267, 170–174. [Google Scholar] [CrossRef]
  199. Fabbri, E.; Rabis, A.; Kötz, R.; Schmidt, T.J. Pt nanoparticles supported on Sb-doped SnO2 porous structures: Developments and issues. Phys. Chem. Chem. Phys. 2014, 16, 13672–13681. [Google Scholar] [CrossRef]
  200. Zaytsev, V.B.; Zhukova, A.A.; Rumyantseva, M.N.; Dobrovolsky, A.A.; Calvo, L.; Gaskov, A.M. Antimony doped whiskers of SnO2 grown from vapor phase. J. Cryst. Growth 2010, 312, 386–390. [Google Scholar] [CrossRef]
  201. Filatova, D.G.; Zhukova, A.A.; Podolko, E.V.; Rumyantseva, M.N.; Gaskov, A.M.; Bolshov, M.A. Determination of antimony and tin in tin dioxide whiskers by inductively coupled plasma mass spectrometry. J. Anal. Chem. 2012, 67, 950–954. [Google Scholar] [CrossRef]
  202. Turkin, A.A.; Chizhov, A.S.; Seregina, I.F.; Filatova, D.G.; Karpov, Y.A. Determination of gold and antimony in advanced materials based on tin dioxide using inductively coupled plasma mass spectrometry. Inorg. Mater. 2015, 51, 1420–1422. [Google Scholar] [CrossRef]
  203. Lopez-Navarrete, E.; Caballero, A.; Orera, V.M.; Lázaro, F.J.; Ocaña, M. Oxidation state and localization of chromium ions in Cr-doped cassiterite and Cr-doped malayaite. Acta Mater. 2003, 51, 2371–2381. [Google Scholar] [CrossRef]
  204. Antony, M.P.; Jha, A.; Tathavadkar, V. Alkali roasting of Indian chromite ores: Thermodynamic and kinetic considerations. Miner. Process. Extr. Metall. 2006, 115, 71–79. [Google Scholar] [CrossRef]
  205. Filatova, D.G.; Rumyantseva, M.N.; Gaskov, A.M. A Method of Determining the Doping Additives of Gold and Cobalt in Semiconductor Materials Based on Tin. Dioxide. Patent RU 2649136, 29 March 2018. [Google Scholar]
  206. Vereda Alonso, E.; García de Torres, A.; Siles Cordero, M.T.; Cano Pavón, J.M. Multivariate optimization of the synthesis and of the microwave dissolution of biomorphic silicon carbide ceramics. Microchem. J. 2011, 97, 101–108. [Google Scholar] [CrossRef]
  207. Filatova, D.G.; Kutukov, P.S.; Rumyantseva, M.N.; Gaskov, A.M. Study of platinum and palladium distribution in advanced materials based on tin dioxide using inductively coupled plasma mass spectrometry (ICP-MS). Inorg. Mater. 2019, 55, 1399–1402. [Google Scholar] [CrossRef]
  208. Klockenkämper, R.; von Bohlen, A. Survey of sampling techniques for solids suitable for microanalysis by total-reflection X-ray fluorescence spectrometry. J. Anal. At. Spectrom. 1999, 14, 571–576. [Google Scholar] [CrossRef]
  209. Hellin, D.; De Gendt, S.; Valckx, N.; Mertens, P.W.; Vinckier, C. Trends in total reflection X-ray fluorescence spectrometry for metallic contamination control in semiconductor nanotechnology. Spectrochim. Acta B 2006, 61, 496–514. [Google Scholar] [CrossRef]
  210. Fernández Ruiz, R. Three empirical cases of the deposition morphology influence in the analytical quality of direct solid suspension measurements by total-reflection X-ray fluorescence. Spectrochim. Acta B 2009, 64, 672–678. [Google Scholar] [CrossRef]
  211. Filatova, D.G.; Alov, N.V.; Marikutsa, A.V.; Seregina, I.F. Ruthenium and palladium determination in advanced materials based on tin dioxide by mass spectrometry with inductively coupled plasma and total reflection X-ray fluorescence. Mosc. Univ. Chem. Bull. 2015, 70, 234–236. [Google Scholar] [CrossRef]
  212. Filatova, D.G.; Alov, N.V.; Sharanov, P.Y.; Marikutsa, A.V. Detecting gold in semiconducting advanced nanomaterials based on tin oxide via total reflection X-ray fluorescence analysis. Mosc. Univ. Chem. Bull. 2015, 70, 60–62. [Google Scholar] [CrossRef]
  213. Filatova, D.G.; Bogdanova, A.P.; Krivetskiy, V.V.; Penkina, T.N.; Rumyantseva, M.N. Quantification of Si dopant in β-Ga2O3-based semiconductor gas sensors by total reflection X-ray fluorescence spectroscopy (TXRF). Ind. Lab. Diagn. Mater. 2022, 88, 5–9. (In Russian) [Google Scholar]
Figure 1. Scheme of the transformation of the phase composition for SnO2–MnOm nanocomposites [45].
Figure 1. Scheme of the transformation of the phase composition for SnO2–MnOm nanocomposites [45].
Materials 16 06733 g001
Table 1. Comparison of characteristics of analysis methods.
Table 1. Comparison of characteristics of analysis methods.
MethodWorking PrincipleLODAnalytical Spot SizeDepth of
Analysis
Information
Energy-dispersive X-ray spectroscopy (EDX)Generation of characteristic X-rays from a specimen through the electron beam0.1–1 at%≥0.1 μm (combined with SEM)
0.05–0.2 nm (combined with STEM)
0.1–3 μmGross elemental composition except for light elements; visualizing element distribution (EDX mapping)
X-ray diffraction (XRD)Interference of monochromatic X-rays ~1% by volume ~20 Å to ~30 µm depending on material properties and X-ray incidence anglePhase composition; determining the crystallographic properties of a sample
Raman spectroscopyScattering of incident light at an energy shifted by the vibrational energy (hν) of the molecule100 ppb~1 µm<10 µmChemical bonding/molecular information; phase composition; surface adsorbates
X-ray photoelectron spectroscopy (XPS)The simultaneous measurement of kinetic energy and the number of electrons escaping when the sample is irradiated with a beam of X-ray radiation under high vacuum0.1–1 at%~10 μm10 nmSurface elemental composition; element oxidation state
Inductively coupled plasma mass spectrometry (ICP MS)Ionization of atoms in plasma and mass selection1–100 pptvolumevolumeDetermining trace elements (<1%) present in a sample
Total reflection X-ray fluorescence (TXRF)Energy-dispersive X-ray fluorescence spectrometry on a thin layer of the sampleng–pgsize of the collimator of the detector (8 mm diameter)volumeGross elemental composition except for light elements; determining trace elements (<1%) present in a sample
High-resolution continuum source graphite furnace atomic absorption spectrometry (HR CS GFAAS)Absorption of ground state-atoms of UV or visible light in the gaseous state10−6–10−1%-volumeDetermining trace elements (<1%) present in a sample
Table 2. Methods of introducing additives that provide the possibility of forming the phase of the second component.
Table 2. Methods of introducing additives that provide the possibility of forming the phase of the second component.
MaterialSynthesis MethodRef.
SnO2/TiO2 nanobeltsI: Hydrothermal (TiO2); II: hydrothermal (SnO2)[136]
SnO2/TiO2 nanotubesI: Electrochemical anodization (TiO2); II: atomic layer deposition (SnO2)[137]
SnO2/TiO2 nanopowdersI: Sol–gel (TiO2); II: sol–gel (SnO2)[138]
TiO2/SnO2 nanosheetsI: Hydrothermal (SnO2); II: pulsed laser deposition (TiO2)[139]
TiO2/SnO2 nanofibersCoaxial electrospinning [140]
TiO2/SnO2 thin filmsI: Magnetron sputtering (SnO2); II: Langmuir–Blodgett (TiO2)[141]
V2O5/SnO2 nanowiresI: Vapor–liquid–solid (SnO2); II: atomic layer deposition (V2O5)[142]
Cr2O3/SnO2 nanofibersCoaxial electrospinning [143]
Cr2O3/SnO2 hollow spheresI: Ultrasonic spray pyrolysis (SnO2); II: E-beam evaporation (Cr2O3)[144]
Cr2O3/SnO2 nanocompositesI: Sol–gel (SnO2); II: electrodeposition (Cr2O3)[145]
Cr2O3/SnO2 nanofibersI: Electrospinning (SnO2); II: impregnation (Cr2O3)[146]
MnO2/SnO2 nanocompositesChemical precipitation with subsequent thermal annealing[147]
SnO2/Mn3O4 nanocompositesI: Plasmo-chemical deposition (Mn3O4); II: RF sputtering (SnO2)[148]
Mn3O4/SnO2 nanocompositesI: Citrate sol–gel (SnO2); II: impregnation (Mn3O4)[149]
Fe2O3/SnO2 core–shellI: Chemical precipitation (SnO2); II: sol–gel (Fe2O3)[150]
Fe2O3/SnO2 hollow spheresI: Hydrothermal (SnO2); II: hydrothermal (Fe2O3)[151]
Fe2O3/SnO2 nanocompositesCitrate sol–gel[152]
Fe2O3/SnO2 nanorodsI: Hydrothermal (SnO2); II: ionic layer adsorption reaction (Fe2O3)[153]
Fe2O3/SnO2 nanofibersI: Electrospinning (SnO2); II: hard-template method (Fe2O3)[154]
Fe2O3/SnO2 nanofibersElectrospinning[155]
Co3O4/SnO2 nanocompositesI: Chemical precipitation (SnO2); II: impregnation (Co3O4)[63]
Co3O4/SnO2 nanowiresI: Vapor–liquid–solid (SnO2); II: DC sputtering (Co3O4)[156]
Co3O4/SnO2 nanowiresI: Vapor–liquid–solid (SnO2); II: sol–gel (Co3O4)[157]
NiO/SnO2 nanowiresI: Vapor–liquid–solid (SnO2); II: atomic layer deposition (NiO)[158]
NiO/SnO2 nanosheetsI: Hydrothermal (SnO2); II: hydrothermal (NiO)[159]
NiO/SnO2 nanocompositesI: Chemical precipitation (SnO2); II: sol–gel (NiO)[160]
NiO/SnO2 hollow spheresI: Hydrothermal (SnO2); II: pulsed laser deposition (NiO)[161]
CuO/SnO2 hollow nanofibersElectrospinning[162]
CuO/SnO2 thin filmsI: RF magnetron sputtering (SnO2); II: RF magnetron sputtering (CuO)[163]
ZnO/SnO2 rootstock/scionI: Vapor–liquid–solid (SnO2 or ZnO); II: vapor–liquid–solid (ZnO or SnO2)[164]
ZnO/SnO2 nanofibersElectrospinning[165]
ZnO/SnO2 nanoheterostructuresI: Hydrothermal (SnO2); II: chemical bath deposition (ZnO)[166]
SnO2/Nb2O5 core–shellI: Hydrothermal (Nb2O5); II: atomic layer deposition (SnO2)[167]
Nb2O5/SnO2 nanocompositesI: Spin coating (SnO2 nanosheets); II: hydrothermal (Nb2O5 nanorods)[168]
Nb2O5/SnO2 nanocompositesI: Chemical precipitation (SnO2); II: impregnation (Nb2O5)[115]
RuO2/SnO2 nanopowdersI: Chemical precipitation (SnO2); II: impregnation (RuO2)[169]
RuO2/SnO2 nanopowdersI: Chemical precipitation (SnO2); II: deposition–precipitation (RuO2)[120]
Table 3. Methods of the synthesis of multicomponent nanocrystalline materials that benefit solid solution formation.
Table 3. Methods of the synthesis of multicomponent nanocrystalline materials that benefit solid solution formation.
MaterialSynthesis MethodRef.
SnO2(Sb) thin filmsSol–gel spin coating[83]
SnO2(Sb) nanopowdersHydrothermal[84]
SnO2(Sb) nanowiresSb-ion implantation[85]
(SnO2:TiO2) thin filmsPulsed laser deposition[87]
(SnO2:TiO2) nanopowdersCo-precipitation[88]
(SnO2:TiO2) nanopowdersFlame spray pyrolysis[89]
(Cr2O3/SnO2) nanopowdersFlame spray pyrolysis [65]
SnO2(Mn) nanobeltsThermal evaporation[91]
SnO2(Mn) thin filmsSpray pyrolysis[92]
SnO2(Mn) nanopowdersCo-precipitation[93]
SnO2(Fe) thin filmsSpray pyrolysis[96]
SnO2(Fe) nanopowdersCo-precipitation, hydrothermal[95]
SnO2(Fe) nanopowdersCo-precipitation, hydrothermal[97]
SnO2(Co) thin filmsSpray pyrolysis[96]
SnO2(Co) thin filmsSpray pyrolysis[100]
SnO2(Co) nanopowdersCo-precipitation, hydrothermal[99]
SnO2(Co) inverse opalUltrasonic nebulizing deposition with a self-assembly template[101]
SnO2(Ni) thin filmsSpray pyrolysis[96]
SnO2(Ni) nanopowdersCo-precipitation, hydrothermal[95]
SnO2(Ni) porous structuresCo-precipitation, hydrothermal[103]
SnO2(Ni) nanorodsCo-precipitation, hydrothermal[104]
SnO2(Ni) nanopowdersCo-precipitation, microwave treatment[105]
SnO2(Cu) porous structuresSurfactant-assisted co-precipitation[107]
SnO2(Cu)/rGO nanocompositesSolvothermal[108]
SnO2(Cu) nanopowdersCo-precipitation[109]
SnO2(Zn) thin filmsSpray pyrolysis[112]
SnO2(Zn) nanostructuresCo-precipitation, hydrothermal[110]
SnO2(Nb) nanopowdersCo-precipitation, hydrothermal[114]
SnO2(Nb) nanopowdersFlame spray pyrolysis[113]
SnO2(Ru) nanopowdersCo-precipitation[120]
SnO2(Ru) nanofibersElectrospinning[117]
SnO2(Ru) nanotubesElectrospinning[118]
SnO2(Ru) thin filmsSpray pyrolysis[119]
Table 4. Examples of reagents and treatment conditions for flushing the additive from the SnO2 surface without affecting the tin dioxide matrix.
Table 4. Examples of reagents and treatment conditions for flushing the additive from the SnO2 surface without affecting the tin dioxide matrix.
Additive on SnO2 SurfaceReagent,
Treatment Conditions
Analysis Method
Sb3+10% citric acid, heating for 30 min in an ultrasonic bath.ICP MS, TXRF
Cr3+1% KSCN in 25% NH3 with the formation of NH4[Cr(NCS)4(NH3)2].ICP MS, TXRF
Co2+, Co3+HNO3:HCl = 1:3.ICP MS, TXRF
PtOx, PdOx, Au0HNO3:HCl = 1:3.ICP MS
RuOxOxalic acid or ascorbic acid (pH = 3), HNO3:HCl = 1:3.ICP MS
MnOxOxalic acid (pH = 3).ICP MS, TXRF
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Filatova, D.; Rumyantseva, M. Additives in Nanocrystalline Tin Dioxide: Recent Progress in the Characterization of Materials for Gas Sensor Applications. Materials 2023, 16, 6733. https://doi.org/10.3390/ma16206733

AMA Style

Filatova D, Rumyantseva M. Additives in Nanocrystalline Tin Dioxide: Recent Progress in the Characterization of Materials for Gas Sensor Applications. Materials. 2023; 16(20):6733. https://doi.org/10.3390/ma16206733

Chicago/Turabian Style

Filatova, Darya, and Marina Rumyantseva. 2023. "Additives in Nanocrystalline Tin Dioxide: Recent Progress in the Characterization of Materials for Gas Sensor Applications" Materials 16, no. 20: 6733. https://doi.org/10.3390/ma16206733

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