Next Article in Journal / Special Issue
Tabulated Chemistry Models for Numerical Simulation of Combustion Flow Field
Previous Article in Journal
Numerical Investigation of Cylindrical Water Droplets Subjected to Air Shock Loading at a High Weber Number
Previous Article in Special Issue
Soot and Flame Structures in Turbulent Partially Premixed Jet Flames of Pre-Evaporated Diesel Surrogates with Admixture of OMEn
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Solution Combustion Synthesis for Various Applications: A Review of the Mixed-Fuel Approach

by
Samantha Padayatchee
1,
Halliru Ibrahim
1,*,
Holger B. Friedrich
2,
Ezra J. Olivier
3 and
Pinkie Ntola
1
1
Department of Chemistry, Durban University of Technology, P.O. Box 1334, Durban 4000, South Africa
2
School of Chemistry and Physics, University of KwaZulu-Natal, Private Bag X54001, Durban 4000, South Africa
3
Centre for HRTEM, Physics Department, Nelson Mandela University, Port Elizabeth 6031, South Africa
*
Author to whom correspondence should be addressed.
Fluids 2025, 10(4), 82; https://doi.org/10.3390/fluids10040082
Submission received: 27 February 2025 / Revised: 21 March 2025 / Accepted: 22 March 2025 / Published: 25 March 2025
(This article belongs to the Special Issue Turbulence and Combustion)

Abstract

:
As solution combustion synthesis (SCS) becomes a universal route to metal oxide nanomaterials, it also paves the way for mixed-fuel combustion synthesis as an advanced approach to the synthesis of materials of desirable properties for diverse applications. Major significance is attached to the rates of decomposition and combustion temperatures of the fuel as determinant factors of the morphology and physicochemical properties of the materials obtained. This has promoted the use of mixed-fuel systems characterized by lower decomposition temperatures of organic fuels and higher rates of combustion. The review work presented herein provides a comprehensive analysis of the applications of mixed-fuel SCS in ceramics, fuel cells, nanocomposite materials, and the recycling of lithium battery materials while taking into consideration the effects of the mixed-fuel system on the physicochemical and morphological properties of those materials, as compared to their analogues prepared via single-fuel SCS.

1. Introduction

Part of the attractiveness of green chemistry is that it is based on a set of principles that reduces or eliminates the use or generation of hazardous substances in the design, manufacture, and application of chemical products [1]. The combined use of such chemical principles led to rapid development of the green chemistry framework that takes into consideration economically viable processes and environmentally benign products. Solution combustion synthesis (SCS) is a versatile method for the production of material powders for diverse applications [2]. It is a green chemistry approach that is gaining wide acceptance as a viable and cost-effective substitute for the initial approaches to the bulk production of metal oxide nanostructured materials and is now a workhorse technique in materials science [3,4,5,6]. The ease of implementation, high throughput, versatility of chemistries, and capacity for the production of high-surface-area powders are among the successes recorded with the SCS approach and make it an outstanding technique among its contemporaries [7]. The technique is relatively novel in its content and various modes of applications. The design and preparation methods play a prime role in determining the structural, surface characteristics and unique properties for the desired application of the final products [8].
The redox reaction between a suitable oxidizer (usually metal nitrates) and an appropriate organic fuel (i.e., sucrose, urea, citric acid, glycine or hydrazide) is highly favoured towards the formation of desired products in SCS. Deganello et al. summarized the process of SCS into three steps, i.e., formation of the combustion mixture, formation of the gel, and then the combustion of the gel [6]. These steps are achieved at a gradient temperature range of 80 °C to >200 °C (Figure 1) and are characterized by solvent evaporation to combustible gases, self-ignition, and combustion.
Most products of SCS are fabricated metal oxides that are of great technological importance in advanced materials for energy generation [9], catalysis [10,11], ceramics [12,13], fuel and solar cells [5,14], batteries, super capacitors, optics [4,15], metal recycling, and photocatalytic [16] and environmental applications [6,17]. So far, the SCS route has been employed to produce a large variety of solid materials such as metals, sulphides, carbides, nitrides, single or complex metal oxides, metal-doped materials, and alloys [18,19,20,21,22]. Despite all these advantages, the SCS is not free of shortcomings that can somewhat limit its applications.
A recent review by Novistkaya et al. [3] has shed light on such issues arising from the use of the SCS technique, like powder agglomeration, less control on powder morphologies, and the presence of leftover organic impurities from the associated incomplete combustion. A variety of factors including the type of fuel, fuel-to-oxidizer ratio, mixture of fuels, amount of water, ρH and combustion temperature, as well as the presence of additives are identified as responsible for such deviations in the morphological and physicochemical properties of the resultant material [11,23]. The nature, quantity, and/or mixture of the fuel used in any SCS plays a central role in the optimization of the material properties. This is because the fuel plays the central roles of a reducer during the combustion of the mixture/formation of the gel substance, a chelating agent, and a microstructural template [5,6,24]. It is then typical of a fuel for SCS to comprise chelating functionalities that can prevent the precipitation of metal ions and help preserve the compositional homogeneity of all the constituents via the formation of strong coordinate bonds [6,18].
There is a wealth of reviews in the literature on the use of single fuels for various applications [18,21,25]. Several studies suggest that the combination of various fuels can potentially be more effective than using a single fuel in terms of having good control of the adiabatic combustion temperature of the reaction mixture, the type and number of gaseous products released, and other related factors [5,6,25,26]. However, there is no published review that comprehensively discusses the usage of the mixed-fuel approach. This review, hence, intends to provide a holistic perspective of the use of the mixed-fuel solution combustion synthetic approach in various applications. This study provides a comprehensive discussion on the positive influence of the mixed fuel on the morphological and physicochemical parameters of the product materials. Consequently, this work provides prospects in the fine adjustment of the mixed-fuel SCS approach for the development of more efficient materials for energy and environmental applications.

2. The Mixed-Fuel Solution Combustion Synthesis Approach

The role of fuel in influencing the final powder properties with regards to the microstructure and morphology has been extensively studied [18,27,28,29,30]. Based on these studies, fuels are classified with reference to the prevailing functional group on the chemical structure of the compound. Thus, amino, carboxylic, hydroxyl, and multifunctional are the four recognized categories of fuels [6,31]. Some common organic molecules which fall within the categories of fuels in SCS are citric acid (C6H8O7), dextrose (C6H12O6), glycine (C2H5NO2), sucrose (C12H22O11), urea (CH4N2O), and hydrazine hydrate (N2H4. H2O).
In this regard, the use of mixed-fuel SCS has shown unique control over the physicochemical properties of the synthesized materials [6,27,28]. This is due to the influence of the mixed-fuel system in improving the performance of the combustion and hence the nature of the products, which cannot be obtained via the single-fuel SCS [32]. The mixed-fuel SCS offers the advantage of control of the reaction atmosphere such that objective materials with desired properties like oxygen vacancy can be obtained [33]. Material properties such as the surface area, particle size distribution, and agglomeration are easily and conveniently fine-tuned by the mixed-fuel system. Typically, the solution combustion reaction occurs by an exothermic reaction of flammable gases (such as NOx, NH3, CO, etc.) that originate from the decomposition of the starting materials (i.e., oxidisers and fuels) [34]. Hence, the combustion behaviour is dependent on the amount and nature of these gases, which can then be tuned by fuel type [23,27,35]. Various fuel mixtures can be used based on factors such as the fuel’s chemical structure and the chelating ability of the fuel. Other key factors include the ability of the fuel to act as a microstructural template, and the role of a fuel’s reducing valency in regulating the maximum temperature and heating rate of a reaction [23]. Notable examples of such mixed-fuel systems employed in SCS for a specific target material are presented in Table 1.
On this basis, a higher specific surface area and smaller particle size in SrFeO3-δ catalysts for the reduction of nitrobenzene was achieved by Naveenkumar et al. [48] through the use of mixed citric acid, glycine, and oxalic acid fuels (Table 1, entry 2). Similarly, (α, β)-SrAl2O4 powders were reported to be directly formed by using a mixture of urea and glycine fuels in work by Ianos et al. [37]. This contrasts with their failed initial attempt in the use of a single fuel (urea or glycine) to synthesize the SrAl2O4. The failure due to the use of a single fuel in the combustion synthesis is attributed to the low amount of energy released during the combustion process [8]. On the other hand, the direct formation of (α, β)-SrAl2O4 powders in the mixed-fuel approach was ascribed to the enhanced combustion performance of mixed fuel via a very substantial increase of released gases, resulting in the spontaneous ignition to a flaming combustion reaction [6,8].
Research findings have shown that glycerine combined with urea or any other amino mixed fuel is best suited for the synthesis of bimetallic materials (Table 1, entries 1–5). Combustion reactions that result in the synthesis of metallic oxide materials are usually carried out with mixed fuels based on citric acid with Cetyltrimethylammonium bromide (CTAB) or glycine or a mixture of the two fuels (Table 1). This might be due to the combustion rate of CTAB or glycine which is beneficial for anisotropic growth of the products [16,49]. Where a heavy metal-doped material is targeted in the SCS; a poly mixed-fuel system, as reported for Sr and Fe- doped barium cobaltite, is applied [5]. These approaches have been successfully employed in the synthesis of materials for various applications, including optics, the battery industry, ceramics, nanocomposites, fuel cells, and pigments.

3. Application of Mixed-Fuel SCS in the Preparation of Desirable Nanomaterials

3.1. Ceramics

It is essential to understand that the fine-tuning of crystallite and particle sizes via combustion parameters contributes to desired material characteristics, thereby making mixed-fuel SCS a versatile method for modifying ceramics to meet specific application requirements [12,45]. The control over the crystal structure of the ceramic material mitigates agglomeration and ensures uniform particle distribution during formation, thereby expediting the standardization of the mechanical and thermal properties of the ceramic materials [21]. A study conducted by Lin et al. on the fabrication and evaluation of γ-LiAlO2 ceramic materials identified a highly promising material for use in tritium breeding applications [50]. The mechanical strength of these breeding blanket ceramics assumes a vital role in maintaining material stability [51], which is significant, particularly in relation to the rate efficiency at which tritium is released and recovered. In this vein, a comprehensive evaluation of the morphology, mechanical properties, and electrical characteristics of γ-LiAlO2 ceramics under a variety of conditions was performed using the mixed-fuel SCS approach.
In another development, the preparation of γ-LiAlO2 involved the use of glycine/urea/leucine–nitrate combinations [52]. An aqueous solution comprising glycine and urea, or glycine and leucine in conjunction with metal nitrates, served as the components (mixed fuel and metal precursor, respectively) for the synthesis of the ceramic. The simultaneous thermal analysis (STA) of precursors obtained under conditions of fuel and oxidizer stoichiometry showed the presence of impurities attributed to incomplete decomposition of initial salts containing carbon fragments of fuel and nitrate groups. An exception was the precursor from the dual-fuel SCS reaction, φ (glycine:urea) = 1:3, in which pure γ-LiAlO2 powder was formed. Particle morphology, phase analysis, and particle size analysis confirmed that a highly reactive pure γ-LiAlO2 with the characteristic fine crystalline texture was successfully synthesized. However, the replacement of lithium nitrate with lithium carbonate was found to allow one to reduce the process temperature and the relative amount of organic fuel. As a result, the content of carbon fragments in the product significantly decreased after synthesis. Under these specific conditions, the γ-LiAlO2 ceramics displayed a consistent microstructure, increased conductivity of the lithium metal, and superior bending strength in comparison to that of the synthesized materials from other fuel mixtures/ratios. These properties aligned with the intended outcomes of the mixed-fuel approach [28]. Remarkably, both the lithium-rich ceramics and those with lithium-deficient phases demonstrated improved ionic conductivities, although it is probable that the mechanisms driving these enhancements are different [50,53].

3.2. Lithium Batteries

The mixed-fuel approach in the SCS of battery materials is pivotal for achieving a high specific discharge capacity and excellent capacity retention [54]. The control offered by SCS on the physicochemical properties of the material allows for optimized ion transport and enhanced reactivity and ensures uniform distribution in the battery materials, with exceptional electrochemical performance for efficient energy storage [9,55].
The phase evolution occurring in the mixed CTAB and glycine fuel during SCS of hierarchical porous LiFePO4 powders was found to be dependent on the nature of the intermediate phases [56]. Thus, the combustion synthesis using an appropriate amount of the mixed fuels followed by calcination of the precursors at 700 °C (Scheme 1) led to the production of highly crystalline single-phase LiFePO4 electrode materials with a high specific discharge capacity and capacity retention [56]. The structural and microstructural properties of the powders, including crystallinity, specific surface area, and particle size, were found to be crucial factors in determining their electrochemical performance [57]. The hierarchical porous microstructure and small particle size of the powders are beneficial for the electrode kinetics [58,59], leading to improved desired properties in terms of the charge/discharge capabilities. Additionally, the porous microstructure and small particle size were found to benefit the electrode kinetics, as indicated by electrochemical spectroscopy [60].
In a recent and sustainable development, the use of mixed-fuel SCS as a recycling approach for spent lithium-ion batteries (LIBs) has become a subject of interest to material scientists. The approach has eased the high cost of materials, low yield due to degradation of materials, risk of environmental contamination, and other challenges posed with the use of either the open-loop [61] or the chemical synthetic [7,62] recycling approaches of the cathode material from the leachate of spent LIBs. This recycling approach, which is based on a self-propagating high-temperature process; offers a simple, flexible, low-cost, and high-yielding method for the preparation of the cathode materials of LIBs. Suitable organic fuels for regenerating cathode materials in the SCS include urea, citric acid, and glycine. The pretreatment and recovery process involves the dissolution into the leachate of spent cathode material, after which the precursor solution is gelled and ignited by increasing the prevailing temperature [63]. While burning the organic fuel, the heat and gaseous products are liberated during the combustion reaction, providing the required thermal energy for crystallisation of the final products with a spongy and foamy microstructure [64].
Mahanipour et al. regenerated a cathode material with a high specific surface area and porous morphology (Scheme 2) from mixed-fuel SCS of spent LIBs [9]. At the beginning, the spent cells were soaked in 10% (w/v) brine solution for the removal of residual electric charge and then manually disassembled to allow the physical separation of the cathode electrodes. These were then pulverized into particulate forms and treated with aqueous NaOH solution (2.5 M, pulp density of 100 g/L) at ambient temperature for 2 h to remove the Al current collector and electrolyte. Burning of the pretreated materials at 650 °C for 4 h allowed for the successive removal of the organic solvent and binder. The recovery of the lithium content in the cathode material was selectively achieved by leaching with oxalic acid. Separation of the leached mixture gave a lithium oxalate solution (C2HLiO4·H2O) as the filtrate, while the residue consisted of the dihydrate triplex nickel-cobalt-manganese oxalate (Ni0.6Co0.2Mn0.2)C2O4·2H2O). The removal of all volatiles by heating the filtrate at 600 °C for 5 h crystallized out the lithium carbonate, while the other residual triplex metal oxalate was oxidized to the corresponding oxide by calcination at 650 °C for 4 h under aerobic conditions [53]. Both recovered materials were then transformed to the nitrate precursors by dissolving in 2 M HNO3 with vigorous stirring at high temperatures. The resulting solution was then used in the SCS approach by treatment with citric acid, CTAB, polyvinylpyrrolidone (PVP), and a mixture of CTAB and PVP fuels in a fuel-to-metal-precursor molar ratio of 4:1. The combustion reaction by the various organic fuels was monitored, and the ignition behaviour of the respective gel formed from each set of the fuel mixture was observed (Scheme 2). The afforded combusted products were successively crushed and calcined at 480 °C for 5 h in an air atmosphere and then heat-treated at 750 °C for 15 h under an oxygen atmosphere to yield the regenerated LiNi0.6Co0.2Mn0.2O2 cathode powders from the spent LIBs.
The structural, microstructural, and electrochemical properties of the regenerated LiNi0.6Co0.2Mn0.2O2 cathode powders are assumed to be influenced by the fuel type or mixture involved in the SC recycling. The slow combustion reaction rate of mixed CTAB–PVP fuels led to the better layered crystal structure of the regenerated cathode materials. Furthermore, the liberation of a large amount of gaseous products is credited for the porous structure with high specific surface area in the synthesized material by using a mixture of CTAB and PVP fuels. The CTAB–PVP as-synthesized powders were found to have a higher electro-chemical performance, including a high discharge specific capacity of 155.9 mAh g−1 at 0.1 C and a high-capacity retention rate of 94% following 100 charging/discharging cycles at a current rate of 0.5 C [9]. The outcomes showed that the mixed-fuel SCS method is a facile and simple route for regenerating the cathode materials from the spent LIBs with high crystal quality and good electrochemical properties.

3.3. Pigments

Precise control of stoichiometry is essential for a pigment spinel material as it forms its cubic Fd⁻3m crystal structure. In this context, Chamyani et al. employed various fuel mixtures for the SCS to evaluate the characteristic structural and physicochemical impact on CoCr2O4 ceramic pigment nanoparticles [65]. Mixtures of the fuels including ethylenediamine/oxalic acid (En/Ox); ethylenediamine/citric acid (En/Cit); oxalic acid/citric acid (Ox/Cit); and a triplex mixture of ethylenediamine, oxalic acid, and citric acid (En/Ox/Cit) resulted in the production of greenish-blue cobalt chromite ceramic pigments. However, the synthesized materials from En/Cit and En/Ox/Cit mixed fuels exhibited higher purity, smaller crystallite size, and a lower degree of agglomeration when compared to those from other fuel mixtures.
The synthetic route for the CoCr2O4 ceramic pigment nanoparticles followed a general pattern where the mixture of the aqueous solutions of the stoichiometric amounts of metal nitrates and fuel were first made in the reaction vessel [66]. The resulting mixture was homogenized by vigorous stirring at 90 °C for 2 h. The subsequent increase in the heating temperature to 300 °C resulted in the self-ignition of the solution in the reaction vessel. The ignition reaction occurred spontaneously with the characteristic release of a significant amount of gases and the formation of a fluffy foamy sample. The reaction temperature was maintained at 300 °C to allow for complete combustion. The final synthesized material was then obtained after calcination at 600 °C for 3 h.
Analysis of the morphological properties of the synthesized pigment materials showed that the utilization of Cit, Ox/Cit, En/Cit, and En/Ox/Cit fuels leads to the formation of CoCr2O4 spinel nanoparticles (Table 2, entries 4–7). In contrast, the use of both non-stoichiometric and stoichiometric En/Ox mixed fuels fails to produce pure CoCr2O4 spinel (Table 2, entries 1–3). It is then worth noting that the choice of fuel types significantly influences both the enthalpy of combustion and the resulting adiabatic flame temperature. The highest adiabatic temperatures were recorded for the as-synthesized samples from entries 6 and 7 in Table 2. The calculated adiabatic temperatures reveal that the Ox/Cit and En/Ox/Cit fuel combinations yield the lowest and highest adiabatic temperatures, respectively. Furthermore, the summarized data in Table 2 on the Field-Emission Scanning Electron Microscopy (FESEM) and Energy-Dispersive X-ray (EDX) analyses of the as-synthesized samples from entries 1 and 2 show the broadest particle size distribution and, hence, the highest degree of agglomeration [65]. On the other hand, the as-synthesized samples from entries 6 and 7 displayed the narrowest particle size distribution and, of course, the lowest degree of agglomeration [67]. Additionally, the EDX data revealed that the molar ratio of Cr to Co is approximately 2 for entries 4–7; signifying the presence of a pure CoCr2O4 spinel phase. Further analyses of the synthesized materials’ colour properties with the CIELab colourimetric system and comparison with the values obtained with the reference standard CoCr2O4 revealed that the parameter attribute a* of the as-synthesized samples from entries 6 and 7 (in Table 2) are significantly more negative than what has been previously reported for analogous as-synthesized pigments from single-fuel SCS [68,69]. This suggests that the intensity of the green color in the as-synthesized samples 6 and 7 is notably higher. Moreover, the L* values for as-synthesized samples 6 and 7 are lower, suggesting that these samples possess a deeper and darker green hue. In general, the as-synthesized CoCr2O4 pigment materials (Table 2; entries 6 and 7) have the smallest particle size distribution and offer better dark green color applications when compared to their previously reported analogues [69]. The analyzed properties further reinforce that the En/Cit and En/Ox/Cit mixed-fuel SCSs produce spinel nanoparticles that exhibit the desired properties which are only attainable via the mixed-fuel approach.
The mixed-fuel SCS approach in pigment preparation profoundly influences critical properties in the material product. Precise stoichiometry control forms the cubic Fd⁻3m crystal structure, essential for spinel materials. The combustion process mitigates agglomeration, ensuring uniform dispersion and optimal colour, while fine-tuning crystallite and particle size customizes the optical attributes of the material. This approach enables high-purity, tailored compositions for diverse colour applications. In this intricate interplay, solution combustion synthesis with the mixed-fuel approach crafts pigments that meet specific colour and optical criteria.

3.4. Fuel Cells

To improve the performance of Solid Oxide Fuel Cell (SOFC) cathodes, it is crucial to design the final microstructure in a way that promotes efficient gas flow and maximizes the availability of reactive sites [70,71]. In this regard, novel fuel mixtures comprising urea and sucrose at different ratios were investigated for the synthesis of Lanthanum Strontium Manganite (LSM) perovskite, with particular emphasis on the powder’s morphology [27]. The fuel systems featuring both urea and sucrose were mixed in the precursor solution, in a reducers-to-oxidizers stoichiometric ratio of 1:2 for the fuel urea and 1:1 for the fuel sucrose. Both fuel systems resulted in the formation of LSM orthorhombic perovskite, which after calcination at 750 °C yielded the corresponding rhombohedral perovskite La0.9Sr0.1MnO3 powder (Scheme 3). The comparisons of the characterisation results from the synthesized samples with those from the CS of either single fuel indicate that the incorporation of sucrose in the precursor solution in conjunction with urea ensued in an increase in the specific surface area on the synthesized rhombohedral perovskite [27]. The mixed-fuel system modified the agglomerates’ microstructure, which tend to form a wide interconnected sub-micrometric porous structure, while the single use of sucrose led to the formation of very sparse agglomerates with an extremely thin microstructure [27]. The former facilitates gas flow and enhances the reactivity of the cathode materials [14,72].

3.5. Nanocomposites

Low oxidation state metal compounds, such as metal nitrates and chlorides, are chosen for their solubility and reduction capabilities in the formation of nanoparticles for various applications [73]. The phases that emerge during the synthesis determine the eventual structure, often consisting of nanoparticles or nanocrystals embedded within a matrix material. The mixed-fuel SCS approach, which is characterized by the combination of two or more fuels and an oxidizer, provides fine-tuned control over the synthesis, influencing nanoparticle formation and, of course, the matrix material’s composition and structure [35]. In a related study conducted by Aruna et al. [44], stable, low-oxidation-state nanocomposites of CeO2-CeAlO3 powders were produced using the solution combustion method. The study utilized two distinct fuel combinations viz.: (a) solely urea and (b) a combination of urea and glycine, in conjunction with the respective metal nitrates as precursors. When urea was used as the only fuel, the resultant product comprised a nano-crystalline CeO2 phase, while Al2O3 was present in an amorphous state. On the other hand, when the mixture of the fuels was employed, a combination of nano-sized CeO2 and CeAlO3 phases was obtained.
The use of a mixture of fuels thus favoured the formation of CeAlO3 and resulted in smaller crystallites, leading to enhanced sintering capabilities as compared to the products obtained from the combustion of a single fuel (urea). Characterisation of the samples from either fuel systems using images from Transmission Electron Microscopy (TEM) revealed that particles of the sample prepared with the mixed fuel displayed a well-packed microstructure composed of a narrow size distribution, with an average size of approximately 10 nanometers (Figure 2b). In contrast, the particles of the sample obtained from using solely urea as fuel exhibited a connected but porous network of larger grains (Figure 2a). Similarly, the Selected Area Electron Diffraction (SAED) pattern obtained from the solely urea synthesized sample is consistent with its X-ray diffraction result, and all the observed diffraction rings index to the CeO2 phase (Figure 2a). However, the corresponding SAED pattern of the sample synthesized from a combination of urea and glycine (Figure 2b) showed rings due to both CeAlO3 and CeO2. All the observed rings are easily indexed to either of the phases. Though overlap of the two phases exists, the appearance of some rings exclusive to each of the phases (i.e., rings shown for CeO2 at (1 1 1) and (0 2 4), while for CeAlO3 at (1 1 1)) [44] confirmed that the synthesized sample from the mixed-fuel approach is indeed a nano-composite of CeAlO3 and CeO2.
The smaller crystallite sizes achieved in the mixed-fuel synthesized sample contribute to improved sintering compared to that of the solely urea sample, which has relatively larger particles. This difference in microstructure of the sintered pellets directly explains the observed variation in the densities of the sintered materials. The combustion synthesis method using a mixture of fuels is hence suggested as a viable route to stabilize low-oxidation-state compounds, such as CeAlO3, which are highly desirable for the development of nanocomposites. This approach can enhance the sintering process and lead to the formation of materials with improved properties and characteristics.

3.6. Dielectrics

Lanthanum aluminate (LaAlO3), known for its perovskite-type structure [74], has gained significant attention due to its diverse range of applications, including its utility in dielectric resonators [75,76]. The traditional solid-state reaction process for the synthesis of the LaAlO3 perovskite presents inherent challenges, such as the need for high reaction temperatures, the formation of large crystallite sizes, limited chemical uniformity, and poor sintering properties [77,78,79]. To address these issues, a new approach that employs the use of a mixed-fuel system of citric acid and oxalic acid as the fuel source along with the corresponding metal nitrates was used to successfully produced nanosized LaAlO3 powders [39]. The synthesis involved the initial mixing of a 1:1 molar ratio of aqueous solutions of the metal precursors. The mix for the fuel systems is meant to have a total fuel concentration of 0.01 M. This allows for the synthesized samples to be described based on the molar ratios of the mixed fuel. Hence, lanthanum aluminate samples LaAlO3-1, LaAlO3-2, and LaAlO3-3 were prepared using mixtures of citric acid and oxalic acid in the molar ratios of 1:3, 1:1, and 3:1, respectively. In each instance, the mixed fuels were added to a La–Al (1:1 molar) aqueous solution and stirred vigorously, while keeping the initial fuel/oxidant molar ratio in the reaction mixture at unity. The resultant sol was continuously stirred for several hours at a constant temperature of 60 °C until it turned into a yellowish sol, after which the temperature was rapidly raised to 80 °C (Scheme 4). This led to visible changes in the viscosity and colour of the sol as it turns into a transparent thick gel. Further heating at the elevated temperature of 200 °C for 2 h turned the gel into a fluffy, polymeric citrate precursor, which was then ground into a powder and calcined at varying furnace temperatures for 2 h to yield the corresponding lanthanum aluminate samples.
The rate of crystallisation of the aluminate samples is influenced by the prevailing temperature and the heating schedule. The X-ray diffraction (XRD) results highlight the significant influence of the fuel mixture ratio on the crystallite size of the synthesized powders. Samples calcined at 700 °C gave a broad continuum XRD pattern (Figure 3i) characteristic of an amorphous sample. Those calcined at 750 °C or 800 °C gave XRD patterns (Figure 3ii,iii) confirming the presence of rhombohedral LaAlO3 with a perovskite structure [39]. Specifically, the calculated crystallite sizes for LaAlO3-1, LaAlO3-2, and LaAlO3-3 were 23.6, 32.2, and 26.8 nm, respectively. These measurements were made for samples calcined at 750 °C and are quite significant when compared to those reported using single-fuel systems of citric acid, oxalic acid, or tartaric acid [24]. In a related development, a suite of pure LaAlO3 nanoparticles prepared from a citrate precursor was reported to have an average range of crystallite sizes of 29–35 nm [80].
It is worth noting that the choice of a mixture of fuels (citric acid and oxalic acid) facilitated a reduction in crystallite size when compared to the product obtained with a single-fuel system. Furthermore, the ratio of the fuel mixture significantly influenced the morphology and size of the resulting powders. LaAlO3-1, prepared using citric acid and oxalic acid (1:3 molar ratio), exhibits the smallest crystallite size of 23.6 nm. This correlates with the average particle size of approximately 41 nm for LaAlO3-1, as determined by Transmission Electron Microscopy (TEM). In contrast, the particles of both LaAlO3-2 and LaAlO3-3 displayed considerable aggregation [39]. Thus, it can be concluded that the utilization of a mixture of fuels not only reduces the exothermicity of the combustion reaction but can also significantly reduce the crystallite size [3,81].

3.7. Optics

The specific material phases generated through mixed-fuel solution combustion synthesis exert a profound influence on the resultant material’s band gap and emissions, thus bearing direct relevance to its optical properties [82]. These multifaceted influences stem from the interplay of various factors such as the crystal structure, phase purity, dopants and defects, quantum confinement effects, and the composition of the synthesized material [15]. Studies on zinc aluminate spinel (ZnAl2O4), have garnered significant attention due to its wide range of applications in sensors, ultraviolet (UV) photo electronic devices, and as catalyst support [4,83,84]. This interest is driven by the unique characteristics of the nanostructured ZnAl2O4 powders, which consist of a broad band gap of approximately 3.8 eV, high fluorescence efficiency, strong chemical and thermal stability, and low surface acidity [85]. In this vein, the mixed-fuel SCS offers the best route to producing zinc aluminate spinel with optimum desirable properties for the various optical applications. The recent work by Mirbagheri et al. utilized a set of molar ratios of urea-glycine mixed fuels for the combustion synthesis of zinc aluminate spinel [4]. The resulting powders were subjected to morphological characterisation, specific surface area measurement, and optical property determination. The findings from the analyses revealed that nearly single-phase ZnAl2O4 powders are obtainable after calcination at 700 °C for the synthesized sample procured in a closed system, whereas intermediate phases are still detectable in some open and closed systems calcined at 600 °C (Figure 4) [4].
The properties of the synthesized zinc aluminate powder are also influenced by the mixed-fuel-to-oxidant ratio (φ). Notably, ZnAl2O4 powders synthesized at a φ value of 0.5 contained both ZnAl2O4 and ZnO phases (Figure 4). The presence of ZnO in the sample via the open system decreased with an increase in the φ value to 2 and resulted in nearly single-phase ZnAl2O4 powders similar to those obtained using the closed system approach (Figure 4i). The reflections due to the ZnO phase in the open system are easily removed with calcination at 700 °C, while powders prepared by the closed system need a lower calcination temperature for complete the formation of the spinel phase (Figure 4). The higher calcination temperature required by powders prepared via the open system approach is attributed to the segregation of intermediate phases formed during the combustion reaction [86]. Furthermore, the existence of the ZnAl2O4 as a combustion product prior to calcination in the open system gave rise to larger-crystallite-size ZnAl2O4 powders when compared to those produced from the closed system. The ZnAl2O4 powders synthesized via the closed system displayed a nearly normal spinel structure, where Zn2+ cations predominantly occupied tetrahedral sites and Al3+ cations were situated in octahedral sites [87]. This normal spinel structure resulted in the absence of blue emissions in the range of 400−500 nm [4,88]. In contrast, the powders prepared in the open system exhibited a smaller band gap energy, measuring 3.52 eV, in comparison to the closed system, which had a band gap energy of 3.90 eV [4].

3.8. Catalyst Supports

The mixed (urea and glycine)-fuel SCS of γ-alumina has proven to be a versatile method for engineering superior catalytic materials [28]. Nanocrystalline γ-alumina, a polymorph of alumina characterized by small crystalline size, has major applications in catalysis and as a support material due to its remarkable attributes like a high surface area and porosity [89,90]. Sherikar et al. initiated research that involved altering the exothermic redox reaction between aluminum nitrate nonahydrate [Al(NO3)3·9H2O] and urea [CO(NH2)2] by using a fuel mixture in the combustion synthesis [28]. In this setup, CO(NH2)2 served as the stoichiometric fuel, and glycine (C2H5NO2) was added as excess fuel. The composition of the fuel mixture was designed so that the amount of urea remained constant under stoichiometric conditions, while glycine was added in multiples of 10 weight percentages, ranging from 0 to 40w% of the urea content to afford five sample compositions that were made with notations U00G, U10G, U20G, U30G, and U40G accordingly (Scheme 5).
X-ray diffraction (XRD) analysis of the synthesized γ-alumina powders revealed interesting phase changes. For instance, with the gradient increase in the glycine content in the fuel, there was a corresponding phase transition from pure α-alumina to a mixture of α- and γ-alumina (Figure 5) [28]. This shift in phases, from the high-temperature alpha phase to the low-temperature gamma phase, is consistent even at theoretically high temperatures and enthalpy and was attributed to incomplete combustion resulting from the high carbon content in glycine [91]. Furthermore, the TEM micrographs displayed the largest aggregate of crystallites with dimensions of 300 nm and a crystallite size of 81 nm (Figure 6) at Ψ = 0 (corresponding to stoichiometric conditions). This is due to the sintering of fine particles at elevated local temperatures [44,59]. However, the crystallite size decreases with the gradient additions of glycine into the fuel composition (i.e., Ψ = 0.4, 0.8, 1.2, 1.6) [28], which lowers the prevailing reaction temperature and consequently prevents particle sintering. The mixed-fuel approach offers a custom-tailored γ-alumina structure with optimized morphology, uniformity, and specific surface characteristics, making it ideally suited for catalytic applications where these properties play a crucial role in achieving high catalytic activity and selectivity [92,93,94].

4. Effect of Mixed Fuels on Combustion of Precursors and Physicochemical and Morphological Properties of the Synthesized Materials

This section discusses the effect of mixed fuels on the combustion reaction of precursors in SCS and how the approach influences different physical and chemical properties, such as particle size, surface area, thermal decomposition temperature and rate, morphology, and phase formation in the synthesized materials.

4.1. Adiabatic Flame Temperature of Combustion Reaction

As discussed earlier in this review, the three prevailing thermodynamic processes that occur during combustion synthesis are heat generation due to the exothermic reaction, gas evolution, and powder crystallisation. The overall effect of these aforementioned processes is on the size of the resulting powders. For instance, high combustion temperatures can result in the formation of relatively large crystallites due to crystal growth, while in some instances, low temperature can prevent particle agglomeration and sintering [48,92]. This analogy is based on the ability of the fuel to increase the combustion temperature and enhance the sample crystallinity [37].
The studies on mixed-fuel combustion synthesis of diverse materials for different applications have demonstrated the effect of mixing fuels on reaction parameters such as adiabatic flame temperatures, maximum temperature reached, and the type of flame formed [28,95,96]. Sherikar and Umarji [28] prepared nano-crystalline γ-alumina powders using a mixture of urea and glycine fuel as reducers. The result for the theoretically calculated variation in adiabatic flame temperature as correlated with the measured flame temperatures demonstrated a direct proportionate relationship between the theoretical adiabatic temperature and the gradient addition of glycine to a maximum of 2020 K (at Ψ = 1.6). In contrast, an inverse relation is observed with the measured flame temperature [28]. The difference in temperature is thus attributed to more moles of gases forming during combustion, which consequently lead to reduced exothermicity of the reaction and incomplete combustion. Similarly, Gotoh et al. demonstrated that different phases of persistent phosphor Gd3Al2Ga3O12:Ce3+-Cr3+ (GAGG:Ce3+-Cr3+) material are formed when single glycine fuel or mixed fuels of glycine and urea at 773 K are used in a solution combustion synthesis [29]. In those reactions’ conditions, a pure phase of GAGG: Ce3+-Cr3+ is obtained with the mixed glycine and urea fuel, while the use of either single fuel (glycine or urea) leads to incomplete combustion and amorphous products.
Sharma et al. also prepared alumina powders using a similar nitrate metal precursor and mixed-urea–glycine-fuel SCS approach [97]. Their study highlighted a continuous change in the various characterisations of the powders with a corresponding change in the percentage combination of the two fuels. When the percentage of urea was higher than that of glycine, the reaction took place with high exothermicity and gave high crystallinity to the product phase, while amorphous character was observed in the cases when higher amounts of glycine were used. The high rate of combustion of urea fuel promotes the agglomeration of particles, while the low combustion of glycine fuel promotes the formation of a small range of particles sizes [28]. Several other authors reported on the effect of exothermicity in the SCS reaction when the glycine was used in a fuel mixture [43,48,91]. In all these studies, the ability of glycine to provide a lower rate of combustion was clearly demonstrated. Such systems are prone to incomplete combustion or combustion with residual carbon content in the product phase. In contrast, systems that are based on increasing urea in the mixed-fuel approach increase the extent of exothermicity in solution combustion [97].
In relation to the effect of mixed fuel on the adiabatic condition for solution combustion synthesis, Khaliullin et al. compiled a list of the temperature effects due the urea/glycine mixture as presented in Table 3 [98]. It follows that some nitrate metal precursors cannot undergo a redox reaction with any fuel. For instance, a mixture of calcium nitrate with urea might not be ignited altogether, while an explosive reaction could be established with β-alanine [99].

4.2. Thermal Decomposition Rate and Temperature

Fuels such as CTAB have higher decomposition temperatures and slower combustion rates than those of glycine and citric acid [100]. However, the combination of CTAB with other fuels such as citric acid and glycine can decrease the decomposition temperature while increasing the combustion reaction rate [13]. The fast combustion reaction rate observed when glycine is used as a second fuel is ascribed to the exothermic reaction of hypergolic gases such as NOx, NH3, CO, etc. formed during the decomposition of the metal nitrate and the fuel [16].
In a related work by Hamedani et al., the use of glycine as a mixed fuel with CTAB to synthesize CoFe2O4 powders gave higher combustibility due to its lower decomposition temperature as compared to that of CTAB fuel [47]. The mixed fuel showed higher weight loss and exothermic temperature than those of single fuels, indicating a slower decomposition rate. Other studies corroborated the aforementioned and revealed that rates of combustion reactions are higher in the samples via mixed urea and glycine fuel when compared to those from the respective single fuels [37].

4.3. Particle Size

Some fuel systems in combustion synthesis serve as structure-directing templates, making surface properties tuneable in metal oxide nanomaterials for potential applications in catalysis, optoelectronics, and gas sensors [31,101,102]. The use of the mixed-fuel combustion synthesis facilitates the reduction in particle size, when compared to the analogous product formed using a single fuel. This is achieved through control of the flame temperature and the type and amount of released gaseous products [27,48]. Asefi et al. noted that BiFeO3 powders prepared from a mixture of glycine and urea had lower crystallite sizes when compared to the powders prepared using the single fuels [32]. This has been reported in similar studies and is attributed to the lowering of the combustion temperature, the evolution of a large volumes of gas (CO2 and H2O), and a less violent reaction caused by the addition of glycine [28,32,44,48,97].
It was observed in a study on the preparation of hierarchical porous LiFePO4 powders for lithium-ion batteries by Karami et al. [56] that a combination with a fuel such as glycine with a high heat content is expected to increase the combustion rate. Morphological analysis of the LiFePO4 as-synthesized powders obtained showed that the crystallite size increased with the content of CTAB in the fuel, while the opposite was observed for the mixed CTAB-glycine. On the other hand, Lazarova et al. used a mixture of glycine–glycerol to prepare nano-sized spinel manganese ferrites [MnFe2O4] at different reducing power ratios [103]. So far, this is the only reported study whose findings oppose the generalized pattern, especially when glycine is used in a mixed-fuel combustion synthesis. This report shows that the synthesized sample from the fuel with the highest amount of glycine had five-times-larger crystallite sizes than the synthesized samples obtained from other fuel mixtures (Figure 7). Their argument is the only outlier in the literature that suggests glycine decreases the exothermicity of a reaction, thereby preventing agglomeration and sintering of the synthesized material [48,92].

4.4. Surface Area

It is generally accepted that the liberation of large amounts of gaseous products such H2O, CO2, N2, H2, and CO during the combustion reaction leads to the formation of a spongy and porous structure [21,25,104,105]. A recent study showed that mixed fuels made up of urea and glycine for the synthesis of Fe3O4 powders produced foamy and sponge-like materials due to the release of large amount of gases within a shorter combustion time [41]. These characteristics favour the synthesis of porous powders and inhibit particle agglomeration [29,41,56]. However, work by Meza-Trujillo et al. [106,107] refuted the generally acceptable norm, as their use of a mixture of urea and β-alanine could not produce a porous material, likely due to the high temperature attained during the combustion reaction which prompted sintering of the grains. The textural properties were only improved by the addition of oxalic acid, which acted as a heterogeneous and combustion retarding agent [107].
The use of activated carbon as a secondary fuel at different fuel–oxidant ratios in the combustion synthesis of ZnO produced a porous material with excellent morphological property for photocatalytic dye removal [108]. Morphological analysis of the synthesized ZnO revealed that the presence of the secondary fuel had a profound effect on reducing the crystallite size and enhancing the specific surface area. The crystallite size of the synthesized ZnO powders varied from 46 to 26 nm, with the powder prepared at a fuel–oxidant ratio of 1.8, which gave the lowest crystallite size and the high specific surface area of 69 m2/g [108]. The excellent absorption and photocatalytic ability of the powder when applied as dye remover are attributed to the synergistic interaction and interplay of the enhanced surface area and catalytic ability of the material [102].

4.5. Material Phases

Alumina powders via the combustion synthesis of glycine-and-urea mixed fuels exist in both the amorphous and crystalline material phases depending on the ratio of the fuel mixture [97]. Powders prepared with a higher content of urea display increased crystallinity, while samples with a higher content of glycine are generally amorphous. These observations reveal that the ratio of the constituent fuels influences phase changes in the material product. In the case of the alumina powders, urea promotes crystallinity, and glycine promotes amorphous character in the product phase. The formation of amorphous phases is attributed to the lack of sufficient temperature to promote alumina crystallisation [97]. Similarly, an increase in the glycine content led to a phase change to amorphous in the case of the mixed (urea and glycine)-fuel combustion synthesis of alumina [28]. The change in phases from the high-temperature α-corundum phase to the low-temperature γ-phase and amorphous phase was reported to be due to incomplete combustion taking place as a result of the high number of carbon atoms present in glycine [28,95].
Studies on single- and mixed-fuel combustion syntheses of FeCr2O4 powders for application in pigments have revealed that phase changes are also influenced by the constituent mixed fuels used [30,109,110]. In this vein, the use of mixed citric acid and ethylene glycol fuels yielded amorphous phases, while the synthesized powders prepared using the glycine fuel resulted in narrower and higher peak intensities in the XRD patterns of the products due to stronger combustion than that with the urea [109,110]. Subsequent addition of glycine to urea, citric acid, or ethylene glycol ensured higher reactivity in all the corresponding fuel mixtures and resulted in the formation of corresponding highly crystalline FeCr2O4 samples. The arrangement of common organic fuels in order of reactivity follows glycine > urea > citric acid > ethylene glycol. Based on this, it is deducible that the mixed-fuel constituents and the fuel mixing ratio affect surface phases, as evident from the associated XRD patterns and FTIR spectra of the synthesized samples [30]. The synthesis of powders with glycine-containing fuel mixtures led to the strongest combustion reaction resulting in the highest crystallinity and the lowest lattice parameters.

5. Summary, Conclusions, and Future Directions

This paper reviewed the uniqueness of the mixed-fuel approach in the solution combustion synthesis of nanomaterials for various applications. The consideration of the triplex function of the fuel and innovative advantages of SCS over its contemporaries led to the growing interest in mixed-fuel combustion synthesis.
Diverse applications of the nanomaterials obtained from mixed-fuel SCS in catalysis, dielectric, recycling, and recoveries of spent LIBs, energy cells, and optics are highlighted in this review. Meanwhile, the corresponding effect of the mixed-fuel system on the combustion and decomposition rates of the precursors as well as on the materials’ physicochemical and morphological properties are discussed in earnest. The synthetic routes and data from the various characterisation techniques like TEM and XRD are used in analyzing and comparing these properties with those reported for analogous as-synthesized materials from single-fuel SCS.
The characteristics of fuels used in SCS such as the molecular structure, functional groups, decomposition temperature, etc. influence their prevailing combustion features such as the combustion reaction rate, combustion temperature, etc. The use of glycine as a fuel in combustion synthesis is promoted by its high combustion rate and temperature and low chelating ability and, hence, results in the low agglomeration and smaller crystallite sizes in the corresponding as-synthesized materials. High-molecular-weight fuels such as CTAB have slow combustion rates and, hence, mixing them with glycine, for instance, increases the combustion rate. Polyfunctional fuels in mixtures have a greater tendency to increase the number of gases evolved during combustion. Thus, the higher decomposition temperature and slow decomposition rate in CTAB made it comparatively unsuitable for combination with citric acid, which is characterized as having high adiabatic temperature due to the low rate of decomposition. In relation to this, common organic fuels follow the order of reactivity glycine > urea > citric acid > ethylene glycol. The synthesis of powders using glycine as fuel led to the strongest reaction, resulting in the highest crystallinity and the lowest lattice parameters. It is then deducible that the nature of the mixed-fuel constituents and the fuel mixing ratio affect XRD patterns, FTIR spectra, and other characterisation spectra of the resultant samples. Hence, with the current emphasis on more environmentally sustainable and benign synthetic routes to chemicals and materials for diverse applications, the prospects are infinite for the usage of mixed-fuel SCS.
The mixed-fuel SC synthetic approach is a viable tool in material chemistry as a cost-effective alternative for the initial approaches to the bulk production of metal oxide nanostructured materials. Ease of implementation, high throughput, versatility of chemistries, and capacity to produce high-surface-area powder materials are among the advantages associated with the SCS. Mixed-fuel SCS as a novel approach to the synthesis of metal oxide nanomaterials is expected to receive a greater audience and acceptance with the publication of this manuscript.

Author Contributions

Conceptualization, S.P., H.I. and P.N.; methodology, S.P., H.I. and P.N.; software, H.I.; validation, S.P., H.I., H.B.F., E.J.O. and P.N.; formal analysis, S.P., H.I., H.B.F., E.J.O. and P.N.; investigation, S.P., H.I., H.B.F., E.J.O. and P.N.; resources, H.I. and P.N.; data curation, S.P., H.I. and P.N.; writing—original draft preparation, S.P., H.I. and P.N.; writing—review and editing, S.P., H.I., H.B.F., E.J.O. and P.N.; visualization, H.I., H.B.F., E.J.O. and P.N.; supervision, P.N., H.B.F. and E.J.O.; project administration, P.N.; funding acquisition, P.N. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by Durban University of Technology (DUT) and the Department of Higher Education and Training of South Africa (DHET).

Data Availability Statement

Not applicable.

Acknowledgments

H.I. would like to acknowledge Durban University of Technology (DUT)/HANT for funding the postdoctoral fellowship, and FCET Gusau for the leave of absence.

Conflicts of Interest

The authors declare no conflicts of interest. The funding sponsors had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, and in the decision to publish the results.

Abbreviations

The following abbreviations are used in this manuscript:
CitCitric acid
CSCombustion Synthesis
CTABCetyltrimethylammonium bromide
EDXEnergy-Dispersive X-ray
EnEthylenediamine
En/CitEthylenediamine/Citric acid
En/OxEthylenediamine/Oxalic acid
En/Ox/CitEthylenediamine, Oxalic acid, and Citric acid
FESEMField-Emission Scanning Electron Microscopy
FTIRFourier Transform Infrared
LIBsLithium-ion Batteries
LSMLanthanum Strontium Manganite
Ox/CitOxalic acid/Citric acid
OxOxalic acid
PVPpolyvinylpyrrolidone
SAEDSelected Area Electron Diffraction
SCSSolution Combustion Synthesis
SOFCSolid Oxide Fuel Cell
STASimultaneous Thermal Analysis
TEMTransmission Electron Microscopy
UVUltraviolet
XRDX-ray Diffraction

References

  1. Anastas, P.T. Introduction:  Green Chemistry. Chem. Rev. 2007, 107, 2167–2168. [Google Scholar] [CrossRef]
  2. Kingsley, J.; Patil, K. A novel combustion process for the synthesis of fine particle α-alumina and related oxide materials. Mater. Lett. 1988, 6, 427–432. [Google Scholar] [CrossRef]
  3. Novitskaya, E.; Kelly, J.P.; Bhaduri, S.; Graeve, O.A. A review of solution combustion synthesis: An analysis of parameters controlling powder characteristics. Int. Mater. Rev. 2021, 66, 188–214. [Google Scholar] [CrossRef]
  4. Mirbagheri, S.; Masoudpanah, S.; Alamolhoda, S. Structural and optical properties of ZnAl2O4 powders synthesized by solution combustion method: Effects of mixture of fuels. Optik 2020, 204, 164170–164177. [Google Scholar] [CrossRef]
  5. Deganello, F.; Liotta, L.F.; Marcì, G.; Fabbri, E.; Traversa, E. Strontium and iron-doped barium cobaltite prepared by solution combustion synthesis: Exploring a mixed-fuel approach for tailored intermediate temperature solid oxide fuel cell cathode materials. Mater. Renew. Sustain. Energy 2013, 2, 8. [Google Scholar] [CrossRef]
  6. Deganello, F.; Tyagi, A.K. Solution combustion synthesis, energy and environment: Best parameters for better materials. Prog Cryst Growth Charact Mater. 2018, 64, 23–61. [Google Scholar] [CrossRef]
  7. Gao, R.; Sun, C.; Zhou, T.; Zhuang, L.; Xie, H. Recycling of LiNi0.5Co0.2Mn0.3O2 Material from Spent Lithium-ion Batteries Using Mixed Organic Acid Leaching and Sol-gel Method. ChemistrySelect 2020, 5, 6482–6490. [Google Scholar] [CrossRef]
  8. Siddique, F.; Gonzalez-Cortes, S.; Mirzaei, A.; Xiao, T.; Rafiq, M.; Zhang, X. Solution combustion synthesis: The relevant metrics for producing advanced and nanostructured photocatalysts. Nanoscale 2022, 14, 11806–11868. [Google Scholar] [CrossRef]
  9. Ghadimi Mahanipour, H.; Masoudpanah, S.M.; Adeli, M.; Hoseini, S.M.H. Regeneration of LiNi0.6Co0.2Mn0.2O2 material from spent lithium-ion batteries by solution combustion synthesis method. J. Energy Storage 2024, 79, 110192–110201. [Google Scholar] [CrossRef]
  10. Masokano, D.S.; Ntola, P.; Mahomed, A.S.; Bala, M.D.; Friedrich, H.B. Influence of support properties on the activity of 2Cr-Fe/MgO-MO2 catalysts (M= Ce, Zr, CeZr and Si) for the dehydrogenation of n-octane with CO2. J. CO2 Util. 2024, 86, 102909–102922. [Google Scholar] [CrossRef]
  11. Ntola, P.; Friedrich, H.B.; Singh, S.; Olivier, E.J.; Farahani, M.; Mahomed, A.S. Effect of the fuel on the surface VOx concentration, speciation and physico-chemical characteristics of solution combustion synthesised VOx/MgO catalysts for n-octane activation. Catal. Comm. 2023, 174, 106571. [Google Scholar] [CrossRef]
  12. Chupakhina, T.I.; Melnikova, N.V.; Kadyrova, N.I.; Mirzorakhimov, A.; Tebenkov, A.V.; Deeva, Y.A.; Zainulin, Y.G.; Gyrdasova, O.I. Synthesis, structure and dielectric properties of new ceramics with K2NiF4-type structure. J. Eur. Ceram. Soc. 2019, 39, 3722–3729. [Google Scholar] [CrossRef]
  13. Masoudpanah, S.; Alamolhoda, S. Solution combustion synthesis of LiMn1.5Ni0.5O4 powders by a mixture of fuels. Ceram. Int. 2019, 45, 22849–22853. [Google Scholar] [CrossRef]
  14. Xu, S.-L.; Zhao, S.; Zeng, W.-J.; Li, S.; Zuo, M.; Lin, Y.; Chu, S.; Chen, P.; Liu, J.; Liang, H.-W. Synthesis of a Hexagonal-Phase Platinum–Lanthanide Alloy as a Durable Fuel-Cell-Cathode Catalyst. Chem. Mater. 2022, 34, 10789–10797. [Google Scholar] [CrossRef]
  15. Zaręba, J.K.; Nyk, M.; Samoć, M. Nonlinear Optical Properties of Emerging Nano-and Microcrystalline Materials. Adv. Opt. Mater. 2021, 9, 2100216–2100250. [Google Scholar] [CrossRef]
  16. Vasei, H.V.; Masoudpanah, S.; Adeli, M.; Aboutalebi, M. Photocatalytic properties of solution combustion synthesized ZnO powders using mixture of CTAB and glycine and citric acid fuels. Adv. Powder. Technol. 2019, 30, 284–291. [Google Scholar] [CrossRef]
  17. Ahmad, I.; Akhtar, M.S.; Ahmed, E.; Ahmad, M.; Keller, V.; Khan, W.Q.; Khalid, N. Rare earth co-doped ZnO photocatalysts: Solution combustion synthesis and environmental applications. Sep. Purif. Technol. 2020, 237, 116328–116340. [Google Scholar] [CrossRef]
  18. Carlos, E.; Martins, R.; Fortunato, E.; Branquinho, R. Solution Combustion Synthesis: Towards a Sustainable Approach for Metal Oxides. Chem. Eur. J. 2020, 26, 9099–9125. [Google Scholar] [CrossRef]
  19. Cao, Z.; Zuo, C. In situ synthesis of chromium carbide nanocomposites from solution combustion synthesis precursors. J. Mol. Struct. 2019, 1175, 496–503. [Google Scholar] [CrossRef]
  20. Roslyakov, S.; Yermekova, Z.; Trusov, G.; Khort, A.; Evdokimenko, N.; Bindiug, D.; Karpenkov, D.; Zhukovskyi, M.; Degtyarenko, A.; Mukasyan, A. One-step solution combustion synthesis of nanostructured transition metal antiperovskite nitride and alloy. Nano Struct. Nano Objects 2021, 28, 100796–100809. [Google Scholar] [CrossRef]
  21. Mukasyan, A.S. Combustion synthesis of boron nitride ceramics: Fundamentals and applications. Nitride Ceram. Combust. Synth. Prop. Appl. 2014, 49–74. [Google Scholar] [CrossRef]
  22. Ramesh, A.; Gavaskar, D.; Nagaraju, P.; Duvvuri, S.; Vanjari, S.R.K.; Subrahmanyam, C. Mn-doped ZnO microspheres prepared by solution combustion synthesis for room temperature NH3 sensing. Appl. Surf. Sci. Adv. 2022, 12, 100349–100360. [Google Scholar] [CrossRef]
  23. Ntola, P.; Friedrich, H.B.; Mahomed, A.S.; Olivier, E.J.; Govender, A.; Singh, S. Exploring the role of fuel on the microstructure of VOx/MgO powders prepared using solution combustion synthesis. Mater. Chem. Phys. 2022, 278, 125602–125613. [Google Scholar] [CrossRef]
  24. Chandradass, J.; Balasubramanian, M.; Kim, K.H. Synthesis and characterization of LaAlO3 nanopowders by various fuels. Mater. Manuf. Process. 2010, 25, 1449–1453. [Google Scholar] [CrossRef]
  25. Varma, A.; Mukasyan, A.S.; Rogachev, A.S.; Manukyan, K.V. Solution combustion synthesis of nanoscale materials. Chem. Rev. 2016, 116, 14493–14586. [Google Scholar] [CrossRef]
  26. Shah, Z.; Veses, R.C.; Brum, J.; Klunk, M.A.; Rocha, L.A.O.; Missaggia, A.B.; Oliveira, A.P.d.; Caetano, N.R. Sewage sludge bio-oil development and characterization. Inventions 2020, 5, 36. [Google Scholar] [CrossRef]
  27. Tarragó, D.P.; Malfatti, C.dF.; de Sousa, V.C. Influence of fuel on morphology of LSM powders obtained by solution combustion synthesis. Powder Technol. 2015, 269, 481–487. [Google Scholar] [CrossRef]
  28. Sherikar, B.N.; Umarji, A.M. Synthesis of γ-alumina by solution combustion method using mixed fuel approach (urea+ glycine fuel). Int. J. Res. Eng. Technol. 2013, 5, 434–438. [Google Scholar] [CrossRef]
  29. Gotoh, T.; Jeem, M.; Zhang, L.; Okinaka, N.; Watanabe, S. Synthesis of yellow persistent phosphor garnet by mixed fuel solution combustion synthesis and its characteristic. J. Phys. Chem. Solids 2020, 142, 109436–109443. [Google Scholar] [CrossRef]
  30. Paborji, F.; Shafiee Afarani, M.; Arabi, A.M.; Ghahari, M. Solution combustion synthesis of FeCr2O4 powders for pigment applications: Effect of fuel type. Int. J. Appl. Ceram. Technol. 2022, 19, 2406–2418. [Google Scholar] [CrossRef]
  31. Shukla, R.; Tyagi, A. Combustion synthesis: A versatile method for functional materials. In Handbook on Synthesis Strategies for Advanced Materials: Volume-I: Techniques and Fundamentals; Springer: Berlin/Heidelberg, Germany, 2021; pp. 51–78. [Google Scholar] [CrossRef]
  32. Asefi, N.; Masoudpanah, S.; Hasheminiasari, M. Photocatalytic performances of BiFeO3 powders synthesized by solution combustion method: The role of mixed fuels. Mater. Chem. Phys. 2019, 228, 168–174. [Google Scholar] [CrossRef]
  33. Li, Q.; Li, Y.-l.; Li, B.; Hao, Y.-j.; Wang, X.-j.; Liu, R.-h.; Ling, Y.; Liu, X.; Li, F.-t. Construction of adjustable dominant {314} facet of Bi5O7I and facet-oxygen vacancy coupling dependent adsorption and photocatalytic activity. Appl. Catal. B Environ. 2021, 289, 120041–120052. [Google Scholar] [CrossRef]
  34. Thoda, O.; Xanthopoulou, G.; Vekinis, G.; Chroneos, A. Review of recent studies on solution combustion synthesis of nanostructured catalysts. Adv. Eng. Mater. 2018, 20, 1800047–1800077. [Google Scholar] [CrossRef]
  35. Khaliullin, S.M.; Koshkina, A. Influence of fuel on phase formation, morphology, electric and dielectric properties of iron oxides obtained by SCS method. Ceram. Int. 2021, 47, 11942–11950. [Google Scholar] [CrossRef]
  36. Ianoş, R.; Lazău, I.; Păcurariu, C.; Barvinschi, P. Fuel mixture approach for solution combustion synthesis of Ca3Al2O6 powders. Cem. Concr. Res. 2009, 39, 566–572. [Google Scholar] [CrossRef]
  37. Ianoş, R.; Istratie, R.; Păcurariu, C.; Lazău, R. Solution combustion synthesis of strontium aluminate, SrAl2O4, powders: Single-fuel versus fuel-mixture approach. Phys. Chem. Chem. Phys. 2016, 18, 1150–1157. [Google Scholar] [CrossRef]
  38. Bai, J.; Liu, J.; Li, C.; Li, G.; Du, Q. Mixture of fuels approach for solution combustion synthesis of nanoscale MgAl2O4 powders. Adv. Powder Technol. 2011, 22, 72–76. [Google Scholar] [CrossRef]
  39. Chandradass, J.; Kim, K.H. Mixture of fuels approach for the solution combustion synthesis of LaAlO3 nanopowders. Adv. Powder Technol. 2010, 21, 100–105. [Google Scholar] [CrossRef]
  40. Aruna, S.; Rajam, K. Mixture of fuels approach for the solution combustion synthesis of Al2O3–ZrO2 nanocomposite. Mater. Res. Bull. 2004, 39, 157–167. [Google Scholar] [CrossRef]
  41. Parnianfar, H.; Masoudpanah, S.; Alamolhoda, S.; Fathi, H. Mixture of fuels for solution combustion synthesis of porous Fe3O4 powders. J. Magn. Magn. Mater. 2017, 432, 24–29. [Google Scholar] [CrossRef]
  42. Hadadian, S.; Masoudpanah, S.; Alamolhoda, S. Solution combustion synthesis of Fe3O4 powders using mixture of CTAB and citric acid fuels. J. Supercond. Nov. Magn. 2019, 32, 353–360. [Google Scholar] [CrossRef]
  43. Kalantari Bolaghi, Z.; Hasheminiasari, M.; Masoudpanah, S.M. Solution combustion synthesis of ZnO powders using mixture of fuels in closed system. Ceram. Int. 2018, 44, 12684–12690. [Google Scholar] [CrossRef]
  44. Aruna, S.; Kini, N.; Rajam, K. Solution combustion synthesis of CeO2–CeAlO3 nano-composites by mixture-of-fuels approach. Mater. Res. Bull. 2009, 44, 728–733. [Google Scholar] [CrossRef]
  45. Sasikumar, S.; Vijayaraghavan, R. Solution combustion synthesis of bioceramic calcium phosphates by single and mixed fuels—A comparative study. Ceram. Int. 2008, 34, 1373–1379. [Google Scholar] [CrossRef]
  46. Aali, H.; Azizi, N.; Baygi, N.J.; Kermani, F.; Mashreghi, M.; Youssefi, A.; Mollazadeh, S.; Khaki, J.V.; Nasiri, H. High antibacterial and photocatalytic activity of solution combustion synthesized Ni0.5Zn0.5Fe2O4 nanoparticles: Effect of fuel to oxidizer ratio and complex fuels. Ceram. Int. 2019, 45, 19127–19140. [Google Scholar] [CrossRef]
  47. Famenin Nezhad Hamedani, S.; Masoudpanah, S.; Bafghi, M.S.; Asgharinezhad Baloochi, N. Solution combustion synthesis of CoFe2O4 powders using mixture of CTAB and glycine fuels. J. Sol.-Gel. Sci. Technol. 2018, 86, 743–750. [Google Scholar] [CrossRef]
  48. Naveenkumar, A.; Kuruva, P.; Shivakumara, C.; Srilakshmi, C. Mixture of Fuels Approach for the Synthesis of SrFeO3−δ Nanocatalyst and Its Impact on the Catalytic Reduction of Nitrobenzene. Inorg. Chem. 2014, 53, 12178–12185. [Google Scholar] [CrossRef]
  49. Vasei, H.V.; Masoudpanah, S.M.; Adeli, M.; Aboutalebi, M.R. Solution combustion synthesis of ZnO powders using CTAB as fuel. Ceram. Int. 2018, 44, 7741–7745. [Google Scholar] [CrossRef]
  50. Lin, J.; Wen, Z.; Xu, X.; Li, N.; Song, S. Characterization and improvement of water compatibility of γ-LiAlO2 ceramic breeders. Fusion. Eng. Des. 2010, 85, 1162–1166. [Google Scholar] [CrossRef]
  51. Wen, Z.; Gu, Z.; Xu, X.; Zhu, X. Research on the preparation, electrical and mechanical properties of γ-LiAlO2 ceramics. J. Nucl. Mater. 2004, 329, 1283–1286. [Google Scholar] [CrossRef]
  52. Zhuravlev, V.D.; Reznitskikh, O.G.; Ermakova, L.V.; Patrusheva, T.A.; Nefedova, K.V. Simultaneous Thermal Analysis of Lithium Aluminate SCS-Precursors Produced with Different Fuels. Int. J. Self-Propagating High-Temp. Synth. 2023, 32, 208–214. [Google Scholar] [CrossRef]
  53. Sharifidarabad, H.; Zakeri, A.; Adeli, M. Preparation of spinel LiMn2O4 cathode material from used zinc-carbon and lithium-ion batteries. Ceram. Int. 2022, 48, 6663–6671. [Google Scholar] [CrossRef]
  54. Adams, R.A.; Pol, V.G.; Varma, A. Tailored solution combustion synthesis of high performance ZnCo2O4 anode materials for lithium-ion batteries. Ind. Eng. Chem. Res. 2017, 56, 7173–7183. [Google Scholar] [CrossRef]
  55. Chen, M.; Zhang, Y.; Xing, G.; Chou, S.-L.; Tang, Y. Electrochemical energy storage devices working in extreme conditions. Energy Environ. Sci. 2021, 14, 3323–3351. [Google Scholar] [CrossRef]
  56. Karami, M.; Masoudpanah, S.; Rezaie, H. Solution combustion synthesis of hierarchical porous LiFePO4 powders as cathode materials for lithium-ion batteries. Adv. Powder. Technol. 2021, 32, 1935–1942. [Google Scholar] [CrossRef]
  57. Karami, M.; Masoudpanah, S.M.; Rezaei, H.R. Electrochemical Performance of LiFePO4/C Powders Synthesized by Solution Combustion Method as the Lithium-Ion Batteries Cathode Material. J. Adv. Mater. Technol. 2022, 11, 81–93. [Google Scholar] [CrossRef]
  58. Xu, H.; Zhao, Y.; Wang, Q.; He, G.; Chen, H. Supports promote single-atom catalysts toward advanced electrocatalysis. Coord. Chem. Rev. 2022, 451, 214261–214280. [Google Scholar] [CrossRef]
  59. Wu, C.; Wang, Y.; Wu, K.; Zhai, H.; Gong, Z.; Peng, Z.; Jin, S.; Du, Q.; Jiao, K. Sintering kinetics and microstructure analysis of composite mixed ionic and electronic conducting electrodes. Int. J. Energy Res. 2022, 46, 8240–8255. [Google Scholar] [CrossRef]
  60. Ivanishchev, A.V.; Bobrikov, I.A.; Ivanishcheva, I.A.; Ivanshina, O.Y. Study of structural and electrochemical characteristics of LiNi0.33Mn0.33Co0.33O2 electrode at lithium content variation. J. Electroanal. Chem. 2018, 821, 140–151. [Google Scholar] [CrossRef]
  61. Sommerville, R.; Zhu, P.; Rajaeifar, M.A.; Heidrich, O.; Goodship, V.; Kendrick, E. A qualitative assessment of lithium ion battery recycling processes. Resour. Conserv. Recycl. 2021, 165, 105219–105230. [Google Scholar] [CrossRef]
  62. Xing, L.; Lin, S.; Yu, J. Novel recycling approach to regenerate a LiNi0.6Co0.2Mn0.2O2 cathode material from spent lithium-ion batteries. Ind. Eng. Chem. Res. 2021, 60, 10303–10311. [Google Scholar] [CrossRef]
  63. Zheng, J.; Zhou, W.; Ma, Y.; Jin, H.; Guo, L. Combustion synthesis of LiNi1/3Co1/3Mn1/3O2 powders with enhanced electrochemical performance in LIBs. J. Alloys Compd. 2015, 635, 207–212. [Google Scholar] [CrossRef]
  64. Li, W.; Yao, L.; Zhang, X.; Lang, W.; Si, J.; Yang, J.; Li, L. The effect of chelating agent on synthesis and electrochemical properties of LiNi0.6Co0.2Mn0.2O2. SN Appl Sci. 2020, 2, 554. [Google Scholar] [CrossRef]
  65. Chamyani, S.; Salehirad, A.; Oroujzadeh, N.; Fateh, D.S. Effect of fuel type on structural and physicochemical properties of solution combustion synthesized CoCr2O4 ceramic pigment nanoparticles. Ceram. Int. 2018, 44, 7754–7760. [Google Scholar] [CrossRef]
  66. Jain, S.; Adiga, K.; Verneker, V.P. A new approach to thermochemical calculations of condensed fuel-oxidizer mixtures. Combust. Flame 1981, 40, 71–79. [Google Scholar] [CrossRef]
  67. Balakrishnan, A.; Pizette, P.; Martin, C.; Joshi, S.; Saha, B. Effect of particle size in aggregated and agglomerated ceramic powders. Acta Mater. 2010, 58, 802–812. [Google Scholar] [CrossRef]
  68. Mestre, S.; Palacios, M.; Agut, P. Solution combustion synthesis of (Co,Fe)Cr2O4 pigments. J. Eur. Ceram. Soc. 2012, 32, 1995–1999. [Google Scholar] [CrossRef]
  69. Gilabert, J.; Palacios, M.D.; Sanz, V.; Mestre, S. Fuel effect on solution combustion synthesis of Co(Cr,Al)2O4 pigments. Bol. Soc. Esp. Ceram. Vidr. 2017, 56, 215–225. [Google Scholar] [CrossRef]
  70. Dwivedi, S. Solid oxide fuel cell: Materials for anode, cathode and electrolyte. Int. J. Hydrogen Energy. 2020, 45, 23988–24013. [Google Scholar] [CrossRef]
  71. Sun, C.; Hui, R.; Roller, J. Cathode materials for solid oxide fuel cells: A review. J. Solid State Electrochem. 2010, 14, 1125–1144. [Google Scholar] [CrossRef]
  72. Yang, X.-Y.; Chen, L.-H.; Li, Y.; Rooke, J.C.; Sanchez, C.; Su, B.-L. Hierarchically porous materials: Synthesis strategies and structure design. Chem. Soc. Rev. 2017, 46, 481–558. [Google Scholar] [CrossRef]
  73. Jin, X.; Lim, J.; Ha, Y.; Kwon, N.H.; Shin, H.; Kim, I.Y.; Lee, N.-S.; Kim, M.H.; Kim, H.; Hwang, S.-J. A critical role of catalyst morphology in low-temperature synthesis of carbon nanotube–transition metal oxide nanocomposite. Nanoscale 2017, 9, 12416–12424. [Google Scholar] [CrossRef]
  74. Rizwan, M.; Gul, S.; Iqbal, T.; Mushtaq, U.; Farooq, M.H.; Farman, M.; Bibi, R.; Ijaz, M. A review on perovskite lanthanum aluminate (LaAlO3), its properties and applications. Mater. Res. Express. 2019, 6, 112001–112031. [Google Scholar] [CrossRef]
  75. Arzeo, M.; Lombardi, F.; Bauch, T. Microwave losses in MgO, LaAlO3, and (La0.3Sr0.7)(Al0.65Ta0.35)O3 dielectrics at low power and in the millikelvin temperature range. Appl. Phys. Lett. 2014, 104, 212601–212605. [Google Scholar] [CrossRef]
  76. Chen, T.-Y.; Pan, R.-Y.; Fung, K.-Z. Effect of divalent dopants on crystal structure and electrical properties of LaAlO3 perovskite. J. Phys. Chem. Solids 2008, 69, 540–546. [Google Scholar] [CrossRef]
  77. Sim, Y.; Yang, I.; Kwon, D.; Ha, J.-M.; Jung, J.C. Preparation of LaAlO3 perovskite catalysts by simple solid-state method for oxidative coupling of methane. Catal. Today 2020, 352, 134–139. [Google Scholar] [CrossRef]
  78. Lee, M.-H.; Jung, W.-S. Facile synthesis of LaAlO3 and Eu (II)-doped LaAlO3 powders by a solid-state reaction. Ceram. Int. 2015, 41, 5561–5567. [Google Scholar] [CrossRef]
  79. Lee, G.; Kim, I.; Yang, I.; Ha, J.-M.; Na, H.B.; Jung, J.C. Effects of the preparation method on the crystallinity and catalytic activity of LaAlO3 perovskites for oxidative coupling of methan. Appl. Surf. Sci. 2018, 429, 55–61. [Google Scholar] [CrossRef]
  80. Yu, H.-F.; Guo, Y.-M. Effects of heating atmosphere on formation of crystalline citrate-derived LaAlO3 nanoparticles. J. Alloys Compd. 2011, 509, 1984–1988. [Google Scholar] [CrossRef]
  81. Ianoş, R.; Borcănescu, S.; Lazău, R. Large surface area ZnAl2O4 powders prepared by a modified combustion technique. Chem. Eng. J. 2014, 240, 260–263. [Google Scholar] [CrossRef]
  82. Patra, A.; Chandaluri, C.G.; Radhakrishnan, T. Optical materials based on molecular nanoparticles. Nanoscale 2012, 4, 343–359. [Google Scholar] [CrossRef]
  83. Strachowski, T.; Grzanka, E.; Mizeracki, J.; Chlanda, A.; Baran, M.; Małek, M.; Niedziałek, M. Microwave-assisted hydrothermal synthesis of zinc-aluminum spinel ZnAl2O4. Materials 2021, 15, 245. [Google Scholar] [CrossRef]
  84. Birhi, D.N.; Laka, A.F. Study of the efficiency of ZnAl2O4 as green nanocatalyst. Inornatus. Biol. Educ. J. 2024, 4, 11–26. [Google Scholar] [CrossRef]
  85. Maity, G.; Maji, P.; Sain, S.; Das, S.; Kar, T.; Pradhan, S. Microstructure, optical and electrical characterizations of nanocrystalline ZnAl2O4 spinel synthesized by mechanical alloying: Effect of sintering on microstructure and properties. Physica. E Low Dimens. Syst. Nanostruct. 2019, 108, 411–420. [Google Scholar] [CrossRef]
  86. Han, M.; Wang, Z.; Xu, Y.; Wu, R.; Jiao, S.; Chen, Y.; Feng, S. Physical properties of MgAl2O4, CoAl2O4, NiAl2O4, CuAl2O4, and ZnAl2O4 spinels synthesized by a solution combustion method. Mater. Chem. Phys. 2018, 215, 251–258. [Google Scholar] [CrossRef]
  87. Tshabalala, K.G. Synthesis and Characterization of Down−Conversion Nanophosphors. Ph.D. Thesis, University of the Free State, Bloemfontein, South Africa, 2014. [Google Scholar]
  88. Jilani, A.; Yahia, I.; Abdel-wahab, M.S.; Al-ghamdi, A.A.; Alhummiany, H. Correction to: Novel Control of the Synthesis and Band Gap of Zinc Aluminate (ZnAl2O4) by Using a DC/RF Sputtering Technique. Silicon 2019, 11, 577. [Google Scholar] [CrossRef]
  89. Busca, G. Structural, Surface, and Catalytic Properties of Aluminas, Advances in Catalysis; Elsevier: Amsterdam, The Netherlands, 2014; pp. 319–404. [Google Scholar] [CrossRef]
  90. Nikoofar, K.; Shahedi, Y.; Chenarboo, F.J. Nano alumina catalytic applications in organic transformations. Mini-Rev. Org. Chem. 2019, 16, 102–110. [Google Scholar] [CrossRef]
  91. Kazemi, H.; Kermani, F.; Mollazadeh, S.; Vahdati Khakhi, J. The significant role of the glycine-nitrate ratio on the physicochemical properties of CoxZn1−xO nanoparticles. Int. J. Appl. Ceram. Technol. 2020, 17, 1852–1868. [Google Scholar] [CrossRef]
  92. Cai, W.; Zhang, S.; Lv, J.; Chen, J.; Yang, J.; Wang, Y.; Guo, X.; Peng, L.; Ding, W.; Chen, Y. Nanotubular gamma alumina with high-energy external surfaces: Synthesis and high performance for catalysis. ACS Catal. 2017, 7, 4083–4092. [Google Scholar] [CrossRef]
  93. Eliassi, A.; Ranjbar, M. Application of novel gamma alumina nano structure for preparation of dimethyl ether from methanol. Int. J. Nanosci. Nanotechnol. 2014, 10, 13–26. [Google Scholar]
  94. Paranjpe, K.Y. Alpha, beta and gamma alumina as a catalyst-A Review. Pharma Innov. J. 2017, 6, 236–238. [Google Scholar]
  95. Sherikar, B.N.; Sahoo, B.; Umarji, A.M. Effect of fuel and fuel to oxidizer ratio in solution combustion synthesis of nanoceramic powders: MgO, CaO and ZnO. Solid State Sci. 2020, 109, 106426–106434. [Google Scholar] [CrossRef]
  96. Behera, P.; Sarkar, R.; Bhattacharyya, S. Nano alumina: A review of the powder synthesis method. InterCeram Int. Ceram. Rev. 2016, 65, 10–16. [Google Scholar] [CrossRef]
  97. Sharma, A.; Rani, A.; Singh, A.; Modi, O.; Gupta, G.K. Synthesis of alumina powder by the urea–glycine–nitrate combustion process: A mixed fuel approach to nanoscale metal oxides. Appl. Nanosci. 2014, 4, 315–323. [Google Scholar] [CrossRef]
  98. Khaliullin, S.M.; Zhuravlev, V.D.; Bamburov, V.G. Solution-combustion synthesis of oxide nanoparticles from nitrate solutions containing glycine and urea: Thermodynamic aspects. Int. J. Self-Propagating High-Temp. Synth. 2016, 25, 139–148. [Google Scholar] [CrossRef]
  99. Patil, K.C.; Aruna, S.T.; Mimani, T. Combustion synthesis: An update. Curr. Opin. Solid State Mater. Sci. 2002, 6, 507–512. [Google Scholar] [CrossRef]
  100. Goworek, J.; Kierys, A.; Gac, W.; Borowka, A.; Kusak, R. Thermal degradation of CTAB in as-synthesized MCM-41. J. Therm. Anal. Calorim. 2009, 96, 375–382. [Google Scholar] [CrossRef]
  101. Mandal, K.; Ghose, S.; Mandal, M.; Majumder, D.; Talukdar, S.; Chakraborty, I.; Panda, S.K. Notes on Useful Materials and Synthesis Through Various Chemical Solution Techniques, Chemical Solution Synthesis for Materials Design and Thin Film Device Applications; Elsevier: Amsterdam, The Netherlands, 2021; pp. 29–78. [Google Scholar] [CrossRef]
  102. Shukla, R.; Dutta, D.P.; Ramkumar, J.; Mandal, B.P.; Tyagi, A.K. Nanocrystalline Functional Oxide Materials. Springer Handb. Nanomater. 2013, 96, 517–552. [Google Scholar] [CrossRef]
  103. Lazarova, T.; Kovacheva, D.; Cherkezova-Zheleva, Z.; Tyuliev, G. Studies of the possibilities to obtain nanosized MnFe2O4 by solution combustion synthesis. Bulg. Chem. Commun. 2017, 49, 217–224, WOS:000418322200039. [Google Scholar]
  104. Evdokimenko, N.; Yermekova, Z.; Roslyakov, S.; Tkachenko, O.; Kapustin, G.; Bindiug, D.; Kustov, A.; Mukasyan, A.S. Sponge-like CoNi catalysts synthesized by combustion of reactive solutions: Stability and performance for CO2 hydrogenation. Materials 2022, 15, 5129. [Google Scholar] [CrossRef]
  105. Klunk, M.A.; Shah, Z.; Caetano, N.R.; Conceição, R.V.; Wander, P.R.; Dasgupta, S.; Das, M. CO2 sequestration by magnesite mineralisation through interaction of Mg-brine and CO2: Integrated laboratory experiments and computerised geochemical modelling. Int. J. Environ. Stud. 2020, 77, 492–509. [Google Scholar] [CrossRef]
  106. Meza-Trujillo, I.; Devred, F.; Gaigneaux, E.M. Production of high surface area mayenite (C12A7) via an assisted solution combustion synthesis (SCS) toward catalytic soot oxidation. Mater. Res. Bull. 2019, 119, 110542–110551. [Google Scholar] [CrossRef]
  107. Meza-Trujillo, I.; Mary, A.; Pietrzyk, P.; Sojka, Z.; Gaigneaux, E.M. Nature and role of Cu (II) species in doped C12A7 catalysts for soot oxidation. Appl. Catal. B Environ. 2022, 316, 121604–121618. [Google Scholar] [CrossRef]
  108. Lutukurthi, D.N.V.V.K.; Dutta, S.; Behara, D.K. Dual role of activated carbon as fuel and template for solution combustion synthesis of porous zinc oxide powders. J. Am. Ceram. Soc. 2021, 104, 4624–4636. [Google Scholar] [CrossRef]
  109. Paborji, F.; Afarani, M.S.; Arabi, A.M.; Ghahari, M. Synthesis of (Fe,Cr)2O3 solid solution pigment powders for ink application. Int. J. Appl. Ceram. Technol. 2023, 20, 1154–1166. [Google Scholar] [CrossRef]
  110. Paborji, F.; Shafiee Afarani, M.; Arabi, A.M.; Ghahari, M. Phase transformation of FeCr2O4 to (Fe,Cr)2O3 solid solution pigment powders: Effect of post-heating temperature. Int. J. Appl. Ceram. Technol. 2023, 20, 281–293. [Google Scholar] [CrossRef]
Figure 1. Illustration of three main steps in solution combustion synthesis.
Figure 1. Illustration of three main steps in solution combustion synthesis.
Fluids 10 00082 g001
Scheme 1. Preparation of single-phase LiFePO4 electrode materials by mixed-fuel SCS [56].
Scheme 1. Preparation of single-phase LiFePO4 electrode materials by mixed-fuel SCS [56].
Fluids 10 00082 sch001
Scheme 2. Descriptive illustration of steps involved in the regeneration of cathode materials from spent LIBs. Reprinted with permission from Mahanipour et al. [9].
Scheme 2. Descriptive illustration of steps involved in the regeneration of cathode materials from spent LIBs. Reprinted with permission from Mahanipour et al. [9].
Fluids 10 00082 sch002
Scheme 3. Isolated flow chart for the mixed-fuel SCS of Lanthanum Strontium Manganite (LSM) perovskite [26].
Scheme 3. Isolated flow chart for the mixed-fuel SCS of Lanthanum Strontium Manganite (LSM) perovskite [26].
Fluids 10 00082 sch003
Figure 2. TEM images (top) and SAED rings patterns (bottom) for (a) the synthesized sample prepared from solely urea, and (b) the synthesized sample prepared from mixed fuel of urea and glycine. Reprinted with permission [44].
Figure 2. TEM images (top) and SAED rings patterns (bottom) for (a) the synthesized sample prepared from solely urea, and (b) the synthesized sample prepared from mixed fuel of urea and glycine. Reprinted with permission [44].
Fluids 10 00082 g002
Scheme 4. Preparation of the lanthanum aluminate perovskites by the mixed-fuel SCS at varying molar ratios of citric acid and oxalic acid fuels [39].
Scheme 4. Preparation of the lanthanum aluminate perovskites by the mixed-fuel SCS at varying molar ratios of citric acid and oxalic acid fuels [39].
Fluids 10 00082 sch004
Figure 3. XRD patterns of lanthanum aluminate as-synthesized powders prepared from the molar ratios of mixed citric acid/oxalic fuel mixtures: (a) LaAlO3-1; (b) LaAlO3-2; and (c) LaAlO3-3 and calcined at respective temperatures of: (i) 700 °C; (ii) 750 °C; and (iii) 800 °C. Reprinted with permission [39].
Figure 3. XRD patterns of lanthanum aluminate as-synthesized powders prepared from the molar ratios of mixed citric acid/oxalic fuel mixtures: (a) LaAlO3-1; (b) LaAlO3-2; and (c) LaAlO3-3 and calcined at respective temperatures of: (i) 700 °C; (ii) 750 °C; and (iii) 800 °C. Reprinted with permission [39].
Fluids 10 00082 g003
Figure 4. (i) XRD patterns of the calcined powders at 600 °C in open (A) and closed (B) systems at various fuel contents (◼: ZnO and ∇: ZnAl2O4). (ii) XRD patterns and Rietveld refinement results of the calcined powders after calcination at 700 °C and prepared by open (A) and closed (B) system at ϕ value of 2. Reprinted with permission [4].
Figure 4. (i) XRD patterns of the calcined powders at 600 °C in open (A) and closed (B) systems at various fuel contents (◼: ZnO and ∇: ZnAl2O4). (ii) XRD patterns and Rietveld refinement results of the calcined powders after calcination at 700 °C and prepared by open (A) and closed (B) system at ϕ value of 2. Reprinted with permission [4].
Fluids 10 00082 g004
Scheme 5. Mixed (urea/glycerine)-fuel SCS of γ-alumina, with varying number of moles of glycine (Ψ) [28].
Scheme 5. Mixed (urea/glycerine)-fuel SCS of γ-alumina, with varying number of moles of glycine (Ψ) [28].
Fluids 10 00082 sch005
Figure 5. XRD patterns of the synthesized alumina samples showing phase transition from pure alpha alumina to a mixture of alpha and gamma alumina with a corresponding gradient increase in the glycine content in the mixed-fuel composition. Reprinted with permission [28].
Figure 5. XRD patterns of the synthesized alumina samples showing phase transition from pure alpha alumina to a mixture of alpha and gamma alumina with a corresponding gradient increase in the glycine content in the mixed-fuel composition. Reprinted with permission [28].
Fluids 10 00082 g005
Figure 6. TEM images of synthesized alumina samples: (a) U00G, (b) U00G, and (c) U20G. Reprinted with permission [28].
Figure 6. TEM images of synthesized alumina samples: (a) U00G, (b) U00G, and (c) U20G. Reprinted with permission [28].
Fluids 10 00082 g006
Figure 7. TEM images of synthesized nano-sized MnFe2O4 using a mixture of fuels: (a) sucrose–urea (3:1), (b) sucrose-urea (1:3), both at 500 °C; (c) glycine–glycerol (3:1) at 400 °C, and (d) glycine–glycerol (1:3) at 500 °C under inert conditions. Reprinted with permission [103].
Figure 7. TEM images of synthesized nano-sized MnFe2O4 using a mixture of fuels: (a) sucrose–urea (3:1), (b) sucrose-urea (1:3), both at 500 °C; (c) glycine–glycerol (3:1) at 400 °C, and (d) glycine–glycerol (1:3) at 500 °C under inert conditions. Reprinted with permission [103].
Fluids 10 00082 g007
Table 1. Common mixed-fuel systems and target materials in SCS.
Table 1. Common mixed-fuel systems and target materials in SCS.
EntryFuel MixtureMaterialReference
1Urea+ β-alanineCa3Al2O6 powders[36]
2Urea and glycineSrAl2O4 powders[37]
3Urea and glycineZnAl2O4 powders [4]
4Urea and glycineMgAl2O4powders[38]
5Urea, glycine, and starch
6Citric and oxalic acidLaAlO3 powders[39]
7Urea and ammonium acetate Al2O3–ZrO2 nanocomposite[40]
8Urea and glycineFe3O4 powders[41]
9CTAB and citric acid Fe3O4 powders[42]
10Glycine and citric acidZnO powders[43]
11CTAB and glycineZnO powders[16]
12Urea and glycineCeO2–CeAlO3 nanocomposites[44]
13Succinic and citric acidBio-ceramic calcium phosphates[45]
14Urea and sucroseSr-doped lanthanum manganite[27]
15Citric acid, cellulose, sucrose polyethylene glycol Sr and Fe-doped barium cobaltite [5]
16Urea and glycineNi0.5Zn0.5Fe2O4 nanoparticles[46]
17Glycine and CTABBiFeO3powders[32]
18Urea and glycineγ-Alumina[28]
19CTAB and glycineCoFe2O4[47]
20Urea and glycineGd3Al2Ga3O12:Ce3+-Cr3[29]
CTAB = Cetyltrimethylammonium bromide
Table 2. Effect of mixed-fuel SCS approach on the morphology of as-synthesized CoCr2O4 ceramic pigment nanoparticles [65].
Table 2. Effect of mixed-fuel SCS approach on the morphology of as-synthesized CoCr2O4 ceramic pigment nanoparticles [65].
EntryMixed FuelParticle Size (nm)MorphologyAgglomerationAtomic Ratio (Cr:Co)Adiabatic Temp/℃
1En/Ox50–350Rod shapedVery high0.004-
2En/Ox(excess)50–150Rod shapedVery high0.14-
3En(excess)/Ox10–30Rod shapedLow1.78-
4Cit10–30CubicLow2.131262.65
5Ox/Cit35–80CubicLow2.31141.47
6En/Cit10–25CubicVery low2.261411.83
7En/Ox Cit10–25CubicVery low2.261839.93
Table 3. Temperature effect ΔTad for adiabatic SCS reactions involving glycine and urea as adopted from Khaliullin et al. [98].
Table 3. Temperature effect ΔTad for adiabatic SCS reactions involving glycine and urea as adopted from Khaliullin et al. [98].
Product ΔTad
K (glyc/urea)
Product ΔTad,
K (glyc/urea)
Product ΔTad
K (glyc/urea)
Li2O2148/1763LiOH2511/2106Li2CO32885/2436
Na2O1502/1199NaOH2063/1722Na2CO32596/2169
K2O1123/857KOH1830/1519K2CO32435/2051
Rb2O1019/762RbOH1703/1410Rb2CO32321/1952
Cs2O885/643CsOH1673/1381Cs2CO32255/1898
BeO3910/3292Be(OH)24043/3422BeCO33969/3349
MgO2977/2484Mg(OH)23219/2710MgCO33358/2825
CaO2488/2053Ca(OH)22779/2331CaCO33121/2627
SrO2165/1766Sr(OH)22636/2182SrCO33052/2561
BaO1954/1561Ba(OH)22401/2004BaCO32928/2456
Al2O33048/2526Al(OH)33273/2731Al2(CO3)33377/2819
Ga2O33632/3055Ga(OH)33897/3292Ga2(CO3)33849/3249
In2O33364/2816In(OH)33603/3029In2(CO3)33533/2967
Sc2O33083/2563Sc(OH)33238/2702Sc2(CO3)33308/2763
Y2O32740/2265Y(OH)33041/2534Y2(CO3)33241/2713
La2O32804/2326La(OH)33162/2651La2(CO3)33405/2863
TiO23043/2571Ti(OH)41696/1358Ti(CO3)22155/1703
ZrO22636/2205Zr(OH)41372/1059Zr(CO3)21821/1415
HfO22546/2130Hf(OH)41173/891Hf(CO3)21629/1252
MnO2679/2226Mn(OH)22963/2481MnCO33110/2614
FeO2690/2245Fe(OH)22942/2467FeCO33069/2578
CoO2816/2367Co(OH)23036/2565CoCO33106/2628
NiO2955/2474Ni(OH)23135/2636NiCO33315/2797
Fe2O33084/2574Fe(OH)33215/2694Fe2(CO3)33236/2709
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

Padayatchee, S.; Ibrahim, H.; Friedrich, H.B.; Olivier, E.J.; Ntola, P. Solution Combustion Synthesis for Various Applications: A Review of the Mixed-Fuel Approach. Fluids 2025, 10, 82. https://doi.org/10.3390/fluids10040082

AMA Style

Padayatchee S, Ibrahim H, Friedrich HB, Olivier EJ, Ntola P. Solution Combustion Synthesis for Various Applications: A Review of the Mixed-Fuel Approach. Fluids. 2025; 10(4):82. https://doi.org/10.3390/fluids10040082

Chicago/Turabian Style

Padayatchee, Samantha, Halliru Ibrahim, Holger B. Friedrich, Ezra J. Olivier, and Pinkie Ntola. 2025. "Solution Combustion Synthesis for Various Applications: A Review of the Mixed-Fuel Approach" Fluids 10, no. 4: 82. https://doi.org/10.3390/fluids10040082

APA Style

Padayatchee, S., Ibrahim, H., Friedrich, H. B., Olivier, E. J., & Ntola, P. (2025). Solution Combustion Synthesis for Various Applications: A Review of the Mixed-Fuel Approach. Fluids, 10(4), 82. https://doi.org/10.3390/fluids10040082

Article Metrics

Back to TopTop