Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Microwave-assisted synthesis of nanomaterials: a green chemistry perspective and sustainability assessment

T. Sajini *a and Jebin Joseph b
aDepartment of Chemistry, St. Berchmans College (Autonomous) Campus, Mahatma Gandhi University, Kottayam, India. E-mail: sajinijebin@sbcollege.ac.in
bDepartment of Botany, St. Berchmans College (Autonomous) Campus, Mahatma Gandhi University, Kottayam, India

Received 12th July 2025 , Accepted 10th August 2025

First published on 3rd September 2025


Abstract

This review critically evaluates microwave-assisted synthesis (MAS) as a sustainable approach for nanomaterial fabrication, addressing environmental concerns associated with conventional methods. MAS offers significant advantages through rapid, uniform heating that reduces energy consumption, reaction times, and hazardous waste generation. By systematically comparing energy, reaction efficiency, waste reduction, selectivity, product uniformity and scalability of MAS with conventional techniques, this review provides a comprehensive framework for sustainable nanomaterial production. The review applies green chemistry metrics and sustainability assessment tools to evaluate the environmental performance and industrial viability of various MAS protocols. Moreover, the integration of MAS with eco-friendly precursors, including plant extracts, biomolecules, and ionic liquids, for synthesising three key nanomaterial classes, such as metal nanoparticles, carbon quantum dots (CQDs), and hybrid nanocomposites was discussed in detail. The discussion further extends to practical applications in catalysis, environmental remediation, energy storage, and biomedical technologies, highlighting how MAS-derived nanomaterials address contemporary sustainability challenges. The review concludes by identifying current challenges and future directions for advancing MAS toward industrial-scale implementation, emphasizing its potential to transform nanomaterial manufacturing into a more environmentally responsible process aligned with circular economy principles.


image file: d5su00584a-p1.tif

T. Sajini

Dr Sajini T. is an Assistant Professor in the Department of Chemistry at St. Berchmans College (Autonomous), Changanassery, Kerala, India. She received her Master's degree in Chemistry from Mahatma Gandhi University, Kottayam, and completed her PhD from the School of Chemical Sciences, Mahatma Gandhi University, Kottayam. Her research focuses on green synthesis of nanomaterials, nanobiochemistry, and environmental nanotechnology, with an emphasis on developing sustainable and biocompatible nanomaterials for catalytic and biological applications. Her research integrates the principles of green chemistry, sustainability assessment, and the development of eco-friendly nanomaterials. She is actively involved in interdisciplinary collaborations and is committed to promoting sustainable innovations in nanoscience.

image file: d5su00584a-p2.tif

Jebin Joseph

Mr Jebin Joseph is an Assistant Professor in the Department of Botany at St. Berchmans College (Autonomous), Changanassery, Kerala, India. His interdisciplinary research bridges soil chemistry, algal ecology, and sustainable nanotechnology. His primary focus lies in the ecological assessment of soil algal communities and their ecological role in agricultural and plantation ecosystems, alongside the green synthesis of nanoparticles using indigenous plant resources for catalytic and antibacterial applications.



Sustainability spotlight

Conventional nanomaterial synthesis methods typically involve excessive energy consumption and toxic chemicals, and generate significant waste. Our review critically examines microwave-assisted synthesis (MAS) as a sustainable alternative that addresses these challenges through rapid, uniform heating mechanisms that substantially reduce energy usage, processing time, and hazardous waste. This work advances sustainability by providing comprehensive green chemistry metrics and assessment frameworks for MAS protocols, particularly for metal nanoparticles, carbon quantum dots, and hybrid nanocomposites with applications in catalysis, remediation, energy storage, and medicine. Our research aligns with UN SDGs 7 (Affordable and Clean Energy), 9 (Industry, Innovation and Infrastructure), and 12 (Responsible Consumption and Production) by promoting energy-efficient manufacturing techniques, cleaner production pathways, and circular economy principles in nanomaterial fabrication.

1 Introduction

Nanomaterials have emerged as a ground-breaking field with the potential to revolutionize various industries ranging from healthcare to energy production.1 Due to the unique properties exhibited by these materials at the nanoscale, the synthesis of nanomaterials has received considerable attention these days.2 Generally, these properties differ considerably from those of their bulk counterparts3,4 and hence, these nanomaterials are extremely promising for use in a wide range of applications including catalysis, sensing, biomedical imaging, drug delivery, environmental remediation and energy storage.5–7 However, the translation of these promising properties into practical applications has been hampered by synthesis challenges, particularly the environmental and economic costs associated with conventional production methods. The critical question facing the field is whether emerging synthesis techniques can deliver on their promises of sustainability while maintaining the precision and scalability required for industrial applications.

Conventional synthesis approaches for nanomaterials have been extensively criticized for their environmental impact and practical limitations.2,8,9 These chemical routes typically require high temperatures and pressures, utilize toxic solvents and reducing agents, and generate hazardous by-products that pose renowned risks to human health and ecosystems.10 A critical examination of the literature reveals that many such approaches fail to provide comprehensive life-cycle assessments or quantitative comparisons of environmental impact, making it difficult to accurately assess the true magnitude of these problems. Furthermore, the economic costs of conventional methods are often inadequately reported, limiting our understanding of the trade-offs between environmental sustainability and commercial viability.

Sustainable and eco-friendly synthesis methods are essential to minimize the environmental impact and promote greener approaches in nanomaterial production, thereby addressing the limitations of conventional synthesis techniques.11 Green and sustainable approaches, such as microwave-assisted synthesis, sonochemical synthesis, plant extract-mediated synthesis, and biomolecule-assisted synthesis, offer eco-friendly alternatives by reducing energy usage, eliminating harmful reagents, and enabling efficient, scalable production.1 These methods promote cleaner and safer nanomaterial fabrication and enhance material properties for applications in medicine, energy storage, catalysis, and environmental remediation.12 The shift towards sustainable synthesis ensures long-term resource efficiency and supports global efforts toward greener technologies and a circular economy.9 Microwave-assisted synthesis (MAS) as an emerging sustainable technique, promises several advantages over conventional methods.12,13 MAS is a process whereby chemical reactions are carried out using microwave irradiation to achieve uniform and rapid heating of the reaction mixture14. There is a growing global interest in this approach for sustainable nanomaterial fabrication.15

Despite growing interest in MAS applications across metallic nanoparticles,10 metal oxides,16,17 quantum dots,18 and carbon-based nanomaterials,19 the literature exhibits significant methodological inconsistencies that compromise the reliability of performance comparisons. Studies frequently lack standardized characterization protocols, employ different metrics for evaluating synthesis success, and rarely provide direct comparisons with conventional methods under equivalent conditions. Additionally, the claims of reduced reaction times and higher yields are not universally supported across all nanomaterial types, and the energy efficiency comparisons with conventional methods often lack rigorous accounting for total system energy consumption. The assertion that MAS enables “milder conditions” requires contextualization, as microwave heating can create localized hot spots and non-uniform heating profiles that may not represent true process improvements.20 Furthermore, the environmental benefits of using “benign” solvents in MAS are undermined when the overall process still requires substantial energy input and specialized equipment with limited lifecycle assessments.21

Therefore, this review addresses these fundamental gaps by systematically evaluating the evidence supporting MAS as a sustainable nanomaterial synthesis approach. We critically analyze the methodological rigor of existing studies, assess the validity of claimed advantages through comparative analysis, and identify where evidence is strong versus where claims may be overstated. Our evaluation framework examines synthesis parameters, characterization methods, and performance metrics across different nanomaterial types while scrutinizing the quality and comparability of reported data. We critically assess green metrics and sustainability claims, examining both the explicit environmental benefits and potential hidden costs of MAS approaches. Finally, we provide a balanced evaluation of MAS scalability challenges and realistic prospects for industrial translation, distinguishing between demonstrated capabilities and speculative projections to guide future research priorities in sustainable nanomaterial synthesis.

2 Fundamental principles of MAS

The underlying principle of microwave-assisted synthesis distinguishes itself through electromagnetic energy delivery within the 0.3–300 GHz spectrum, creating internal heat generation rather than relying on surface-to-core thermal transfer characteristics of traditional methodologies.22 Conventional heating strategies necessitate sequential energy migration through conductive and convective pathways, inherently producing thermal gradients and extended processing durations.23 The mechanism involved in MAS of nanomaterials is depicted in Fig. 1. Polar molecules or ions absorb microwave radiation in the reaction mixture in microwave-assisted synthesis. Due to this absorption, localized heating occurs at the molecular scale and allows for breaking of chemical bonds and chemical reactions to begin.24 This selective heating of reaction components provides a means of controlling reaction conditions, specifically temperature, pressure, and reaction kinetics, with a degree of precision.25
image file: d5su00584a-f1.tif
Fig. 1 Schematic representation of the mechanism of conventional and microwave heating.

Microwave technology promotes simultaneous molecular agitation via dipole oscillation and charged particle migration throughout the entire reaction volume. While this internal energy deposition theoretically achieves homogeneous temperature profiles and accelerated kinetics, practical implementation reveals significant challenges.25 The vessel configuration, reaction scale, and material dielectric characteristics introduce heterogeneous energy absorption patterns that compromise the uniformity assumption, raising concerns about process reproducibility and commercial scalability potential. If uniform heating, a primary claimed advantage of MAS, cannot be reliably achieved, then comparative advantages over conventional methods require reconsideration.

3 Nanomaterial synthesis via MAS

Continued development of MAS techniques has led to the utility of such methods for rapid and efficient fabrication of nanomaterials tailored with required properties for numerous applications. They take advantage of the speciality of microwave irradiation to both speed up chemical reactions and produce nanomaterials in controlled environments. Hydrothermal2,26 and solvothermal27,28 synthesis are among the diverse MAS methods that use high temperature and pressure to foster the nucleation and growth of nanoparticles in aqueous and organic solvents, respectively. Microwave energy is utilized in various nanomaterial synthesis approaches such as sonochemical synthesis, combustion synthesis, solid-state synthesis, and plasma-assisted synthesis.

Microwave irradiation and ultrasonic waves induced chemical reactions in liquid media, also called ‘sonochemical synthesis’, and provide rapid synthesis of nanomaterials with improved yield and control over the size and shape of the particles.29 Microwave-heated combustion synthesis (MHCS) utilizes microwave heating to initiate exothermic reactions between the metal precursor and fuel source to produce nanomaterials with controlled stoichiometry and phase purity at rapid combustion rates.30 However, MHCS is well-known for its rapid reaction rates, and its scalability and control over byproducts remain underexplored. Microwave irradiated solid state synthesis employs microwave irradiation to trigger the solid-state reactions between solid reactants or precursors for synthesizing various complex nanostructures through solid-state transformations at elevated temperatures.31 Furthermore, plasma synthesis techniques employ microwave energy to produce and maintain plasma discharges for synthesizing nanomaterials whose sizes, shapes and surface properties are controlled.32 Plasma synthesis raises sustainability concerns due to its high energy input, which contradicts the green chemistry goals unless renewable energy is employed.

Due to their high reaction efficiency, scalability and product quality, these MAS techniques are promising tools for the sustainable synthesis of nanomaterials for a variety of applications in catalysis, electronics, energy storage or biomedical engineering. The blanket assertion that MAS methods inherently ensure product quality and sustainability overlooks the need for lifecycle assessments and standardized metrics to substantiate these claims. Therefore, a more nuanced analysis, considering both the technical advancements and practical trade-offs of MAS, is essential for framing its role in sustainable nanomaterial development.

Utilizing microwave irradiation, researchers can efficiently synthesize a broad range of nanostructures, including metal nanoparticles, metal oxides, carbon-based quantum dots, and nanocomposites.15,33,34 A detailed discussion about MAS of three major nanomaterials such as metal nanoparticles, carbon quantum dots (CQDs), and nanocomposites is outlined below. Metal and metal oxide nanoparticles are widely studied for their unique electronic, optical, and catalytic properties,34 while CQDs have attracted growing attention due to their excellent fluorescence, water dispersibility, and low toxicity.35 Nanocomposites, which integrate two or more nanoscale components, offer synergistic effects that improve structural and functional performance.36

3.1 Metal nanoparticles

In recent years, the synthesis of metal nanoparticles (MNPs) has attracted considerable attention because of their unique physicochemical properties and applications in various fields.37 Unique techniques for the synthesis of MNPs, including chemical reduction, sol–gel processing, and hydrothermal methods, have several disadvantages, such as long reaction times, high energy consumption, and the adoption of toxic reducing agents.38 Microwave-assisted synthesis is widely employed in the laboratory for producing metal nanoparticles with controlled size, shape, and composition.39 An existing review article on microwave-assisted synthesis of metal nanoparticles (MNPs) discusses the advantages of this method in obtaining high-purity matrices with near-monodispersity and tunable properties.40

The MAS approach appreciably improves reaction kinetics, induces effective nucleation, and reduces undesired side reactions, thus rendering it highly appropriate for the manufacture of noble and non-noble metal nanoparticles,41 along with bimetallic,42 trimetallic43 and metal-based nanocomposites.44 Microwave irradiation offers more control over reaction parameters than traditional heating, cutting processing times from hours to minutes while increasing yield and reproducibility.45 Additionally, MAS supports green chemistry techniques, which frequently use ionic liquids, plant extracts, or bio-based reducing agents to reduce their negative effects on the environment.46

3.1.1 Noble metal nanoparticles (Ag, Au, Pt, Pd, Rh, and Ir). The distinctive optical, electronic, and catalytic characteristics of noble metal nanoparticles, such as silver (Ag), gold (Au), platinum (Pt), palladium (Pd), rhodium (Rh), and iridium (Ir), have led to extensive research into these materials. Compared to traditional methods, MAS improves their size control, crystallinity, and monodispersity.

Silver nanoparticles (AgNPs) are synthesized using a variety of chemical and biochemical techniques.47 Green synthesis of AgNPs, which uses bio-based precursors for their synthesis, has gained more attention recently as a sustainable approach.48 Catalysis, sensing, and antimicrobial applications all make extensive use of the synthesized AgNPs.49 AgNPs with specific morphologies, including spherical particles, triangular nanoplates, nanowires, and dendritic structures, and narrow size distributions can be produced through microwave synthesis.50 Specific research indicates that the shape of nanoparticles and their bioactivity are greatly influenced by reaction time, precursor concentration51 and the nature of reducing agents; for example, stress-induced plant extract shows favourable size and antibacterial properties compared to normal plants.52

Like AgNPs, gold nanoparticles (AuNPs) are rapidly nucleated and grow when exposed to microwave radiation, producing regulated shapes such as spheres, rods, and hexagonal plates.53 To ensure sustainable synthesis with high catalytic and plasmonic properties, reducing agents such as ionic liquids and biological extracts have been employed.54–56

Palladium (Pd) and platinum (Pt) nanoparticles (NPs) are essential for hydrogen storage, catalysis, and fuel cells. Highly dispersed Pt57 and Pd58,59 nanoparticles with superior electrocatalytic performance are produced by microwave-assisted reduction of metal salts using ethylene glycol, polyvinylpyrrolidone (PVP), and ascorbic acid.40 The synthesis of rhodium (Rh) and iridium (Ir) nanoparticles using a microwave enhances their efficiency in hydrogenation reactions and catalytic applications by producing well-defined nanostructures with regulated oxidation states.40 MAS ensures sustainable synthesis of these noble metal nanoparticles with superior catalytic activities and may show high qualities in various applications.

3.1.2 Non-noble metal nanoparticles (Cu, Ni, Co, Fe, Zn, and Mn). Despite being less expensive, non-noble metal nanoparticles need to be carefully stabilized because of their high oxidation and aggregation susceptibility. Stable nanostructures can be created quickly and under control using MAS techniques. Copper nanoparticles (CuNPs),60,61 copper nanowires, nanospheres, and core–shell structures can be effectively produced by modifying the reaction conditions.62 Iron (Fe),63 cobalt (Co),64 nickel (Ni),64,65 zinc oxide (ZnO)13,66 and manganese dioxide (MnO2) nanoparticles (NPs)5 have been synthesised using MAS and they exhibited a wide range of applications. Rapid nucleation and consistent particle growth are made possible by microwave irradiation, which also increases these particles' properties. By fine-tuning microwave parameters, their crystal structure and surface morphology can be improved.
3.1.3 Bimetallic and trimetallic nanoparticles. Compared to their monometallic counterparts, bimetallic and trimetallic nanoparticles have better synergistic qualities.67 MAS improves heterojunction architectures, alloy formation, and core–shell structures of these nanoparticles40 through its regulated heating and consistent energy distribution. In contrast to their monometallic counterparts, bimetallic nanoparticles (BNPs) are nanostructures made of two distinct metal elements that combine the special qualities of both metals to display superior physicochemical properties.68 Improved catalytic activity, optical behavior, magnetic properties, and stability are the results of the two metals working in concert, which makes BNPs extremely valuable in a variety of applications such as energy storage, sensing, medicine, and catalysis.69 An effective, quick, and environmentally friendly method for creating BNPs with exact size control, morphology, and composition tuning is MAS.70,71 Depending on their composition, Ag–Au alloy nanoparticles have adjustable plasmonic and catalytic characteristics.72 Because of their increased electrocatalytic activity, Pt–Pd73 and Pd–Au core–shell structures are being investigated extensively for fuel cell applications and catalysis. Microwave-synthesised Cu–Ag bimetallic nanoparticles have demonstrated encouraging improvements in electrical conductivity and antimicrobial activity.74 Ag–ZnO BNPs were fabricated via a microwave strategy for better antibacterial activity.75 Fig. 2 provides a schematic illustration of how bimetallic Ag–Cu NPs are made from plant extract via MW irradiation.
image file: d5su00584a-f2.tif
Fig. 2 Schematic representation of the formation of bimetallic Ag–Cu NPs from plant extract.

Trimetallic nanoparticles (TNPs) have drawn a lot of interest because, when three distinct metal elements are combined, they produce synergistic effects that improve their physico-chemical characteristics when compared to those of their monometallic and bimetallic counterparts.76 These nanoparticles are ideal for use in energy conversion, environmental remediation, and the biomedical industry because of their exceptional catalytic activity, electronic characteristics, and thermal stability.77 In methylene blue degradation, Au–Pt–Pd nanofluid synthesised via microwave irradiation shows exceptional stability and catalytic efficiency.78 TNPs such as AuPtCu,43 AgRuNi79 and PtPdCu57 were also prepared via MAS strategies, and they were applicable as an anti-bacterial agent, photocatalysts, and stable catalysts, respectively. Through MAS of trimetallic nanocomposites on CQDs, La/Cu/Zr/CQDs80 and Ag/Ti/Zn/CQDs81 have been fabricated and applied for the removal of malachite green dye and solar cells, respectively.

3.1.4 Metal nanocomposites. Metal-based nanocomposites improve the mechanical, thermal, and functional properties of carbon, polymer, or ceramic matrices by integrating nanoparticles into them.65 The unique properties of MAS guarantee that the nanoparticles are uniformly distributed throughout the matrix. High-performance catalysts, sensors, and energy storage nanocomposites are produced when reduced graphene oxide (rGO) is mixed with metal nanoparticles such as rGO–iron oxide,82 BiOBr/rGO,83 rGO–Ag, rGO–SnO2,84 Ag–ZnFe2O4@rGO,85 and rGO–TiO2.86 Microwave-assisted metal-oxide nanocomposites, like MgFe2O4–ZnO,87 chitosan–metal oxide nanocomposite,88 Co3O4–graphene sheet,89 graphene–SnO2,90 LiFePO4/C,28 cellulose/AgCl,91 graphene–ZnO,92 NaFePO4–C,93 MoS2/PANI,94 carbon–zirconium-incorporated CeO2,95 g-C3N4/SnO2,96 CuO/MnO2,5 ZnO/MWCNT,97 gadolinium/cerium oxide,98 CdO–ZnO,99 ZnO/CuO,100 Sb2O3–Ag101 and graphene/CoMoO4 (ref. 102) exhibit remarkable photocatalytic activity and stability, which makes them perfect for solar energy conversion, dye degradation, antimicrobial application, pollutant degradation, supercapacitors, gas sensors, and lithium and sodium ion batteries. Fig. 3 depicts the antibacterial activity of metal nanocomposites of silver nanoparticles decorated on functionalized MWCNTs synthesised via the microwave method.44
image file: d5su00584a-f3.tif
Fig. 3 Schematic representation of the antibacterial activity of Ag–HC/f-MWCNT. Reproduced with permission from ref. 44. Copyright, 2023, Springer.

Microwave-assisted synthesis has undeniably transformed the landscape of metal nanoparticle fabrication by enabling faster, more energy-efficient, and environmentally conscious routes compared to conventional techniques. While the advantages of MAS, such as reduced reaction time, enhanced monodispersity, and shape control, are highlighted for noble and non-noble metals, the sustainability implications of the precursors used (e.g., ethylene glycol60 and polyvinylpyrrolidone40) and the post-synthesis processing steps are not thoroughly assessed.

Moreover, the increasing complexity from mono- to tri-metallic systems introduces new challenges in phase control, alloying behavior, and surface segregation, yet these are seldom critically discussed in the literature. Assertions regarding “synergistic effects” in bimetallic and trimetallic nanoparticles often lack quantitative validation or mechanistic insights. Likewise, while metal nanocomposites are described as having enhanced performance due to uniform dispersion, the scalability of such uniformity, especially in solid supports like graphene or cellulose, remains a significant hurdle. A more nuanced evaluation is warranted, especially in linking synthesis parameters (microwave power, irradiation time, and solvent type) with property tuning and sustainability metrics. Without standardized benchmarking across studies, the real comparative advantage of MAS remains difficult to quantify, particularly when transitioning from laboratory-scale demonstrations to industrial-scale applications.

3.2 Carbon quantum dots

Nearly spherical metal oxide particles comprising carbon atoms known as carbon quantum dots (CQDs) have gained a noticeable amount of attention due to their ease of functionalization and high biocompatibility.103 CQDs are guaranteed to reach their goals through microwave heating as it allows for direct contact instead of surface contact with solvents and other precursors, turning them into dielectrics.104 CQDs are useful for many applications,105 summarized in Table 1, and the following items are commonly needed for their synthesis. Carbon precursors: they include glucose, citric acid or even agricultural waste such as banana peels.106 Reaction medium: solvents like ethylene glycol, water and ionic liquids, which are capable of soaking up energy from the microwave, are used. Reaction conditions: the desired mixture is commonly exposed to a level of microwave radiation between 500 and 900 W for a time span between 2 and 15 minutes. Wrap-around doping with heteroatoms (like phosphorus, nitrogen and sulfur) boosts the optical performance and photoluminescence of CQDs, ensuring a photoluminescence quantity of 98 percent.107 Functionalization also enhances the binding of CQDs with target molecular species, making their potential application better.108
Table 1 CQDs formed via microwave irradiation and their application
Name of CQDs Precursors/materials used Microwave specification Application Reference
Fluorescence CQDs Roasted chickpeas 350 W Detection of Fe3+ ions 112
N, S co-doped CQDs Citric acid and 4-aminobenzenesulfonic acid 800 W, 5 min Fe3+ and hydroquinone detection 107
Amphoteric fluorescence CQDs Sugarcane bagasse (SB) and citric acid 750 W, 135 min Chromium adsorption 113
Fluorescence CQDs Succinic acid (SA) and tris(2-aminoethyl)amine (TAEA) 700 W, 5 min Multi-color fluorescence imaging of cellular media 114
CQDs Hibiscus rosa-sinensis 700 W, 1 min Antibacterial, anti-inflammatory, and wound healing 26
CQD nanocomposite Anchoring carbon quantum dots onto TiO2 CO2 photoreduction 115
N, S doped CQDs Cyclodextrin, melamine, and thiourea 700 W, 5 min Sensor for silver(I) ions 116
Fluorescent carbon quantum dots Mexican mint extract 1200W Fe3+ detection and bio-imaging 117
CQDs Arabica coffee grounds 3–7 min Detection of Fe3+, Pb2+, and Cr3+ 18
CQDs Raw cashew gum 800 W, 30–40 min 118
Lysine-based CQDs Lysine 800 W Luminescent and biocompatible 119
CQDs L-Cystine 800 W, 30–90 s 120
CQDs Acidic linter waste 400 W, 5 min Fluorescence cancer imaging 121
N-CQDs Ammonia solution of xylan 200 W, 10 min Tetracycline detection 109
CQDs Citric acid (monohydrate) and urea 700 W, 150–300 s 110
N-CQDs Glucose water solution with ammonia hydroxide 100 W and 200 W, 60 s Catalytic 122
CQDs Glucose 500 W, 7–11 min Photocatalytic 123
CDs@MIPs Citric acid 750 W, 2 min 30 s Tetracycline detection 124
CQDs Lignocellulosic residues 100 °C for 120 min 125
CQDs A4/B2 polyamidation monomer sets Cu2+ detection 111
CQDs Ziziphus Mauritiana (ZMS) biomass Detection of NH3 126
CQDs Fenugreek seeds 500 W To grow fluorescent protein crystals 127
S, N-CQDs Cabbage and onion 700 W, 4 min Fluorescence sensor for nitazoxanide and hemoglobin 128
CQDs Citric acid 700 W, 3 min Degradation of rhodamine-B 129
CQDs Bamboo tar 180 °C, 15 min Detection of 2,4,6-trinitrophenol 130
Co–NS-CQDs Citric acid and methionine 1.5 min Detection of monosodium glutamate 131
CQDs Citric acid 800 W, 5 min Cellular imaging 132
Fe–CQDs Citric acid monohydrate and urea 750 W, 150 s Detection of diclofenac sodium 133
CQDs L-Arginine 3 min Detection of Fe3+ ions 134
N,S-CDs Citric acid and thiosemicarbazide 1 min Sensor for copper (Cu2+) and etidronate disodium (ETD) 135
CQDs Calotropis gigantea 2 min Bioimaging 136
CQDs Polyethylene glycol 500 W, 60 min 137
CQDs EDTA, sodium thiosulfate, and urea 2 min Sensor for ascorbic acid and riboflavin 35
CQDs Cellulose and carboxymethyl cellulose 750 W, 135 min Chromium absorption 138
CQDs Quince fruit 30 min Cell imaging and As3+ determination 139
CQDs/Ag nanocomposite Glucose and natural orange juice Colorimetric sensor for losartan potassium 140
CQDs@PAFP Starch 700 W, 10 min Removal of uranium 141
N@CQDs Eruca sativa leaves 700 W, 4 min Estimation of modafinil 142
CQDs Orange peels Photosensitizers of TiO2 108
N, S-CDs Sucrose, urea, and thiourea 4 min Natamycin determination 143
N-CQDs Orange juice and urea 10 min Determination of palbociclib 144
CQDs Crude glycerol residue   Sensing Fe3+ and Cu2+ and the photocatalysis of Victoria Blue B dye 145
S-CQDs Reduced glutathione and L-cysteine 2 min Detection of Co2+ and Ni2+ 146
CQDs Citric acid monohydrate and urea Detection of tartrazine 147
NS-CQDs Citric acid, urea, and p-aminobenzenesulfonic acid 800 W, 3–5 min Detection of Fe3+ ions and hydroquinone (HQ) 148
N-CQDs Succinic acid and gallic acid 5 min Detection of methyl orange in saffron 149
Cu–CQDs Ammonium citrate and CuCl2·2H2O Programmed temperature control Histidine detection 150
CQDs Araucaria heterophylla gum extract 4 min Drug detection 151
NS-CQDs Citric acid and L-cysteine 1 min Determination of ascorbic acid 152
CQDs Ascorbic acid, vitamin C, EDTA, and polyethylene glycol 1 h Dopamine detection 153
N-CQDs Citric acid and urea 15 min White-light emissive diodes 154
CQDs Empty fruit bunch 1 min Detection of Cu2+ 155
NP@CQDs EDTA and diammonium hydrogen phosphate 3 min Determination of glutathione 156
ZnPc–CQDs Zinc(II) tetra-amino-phthalocyanine (ZnPc) and citric acid 700 W, 2 min Antibacterial 157


In MAS of CQDs, a bottom-up approach is followed where small organic precursors are thermally decomposed, polymerized, and carbonized to yield nanoscale carbon quantum dots, which is depicted in Fig. 4.109 Polar molecules of the precursor solution can absorb electromagnetic energy when subjected to microwave irradiation, resulting in broad and uniform heating through dipole rotation and ionic conduction.110 This heating causes an accelerated dehydration and polymerization of the carbonaceous precursors to form intermediate polymeric clusters. With the increase in temperature, these clusters then experience carbonization, allowing the nucleation and growth of graphitic sp2 carbon domains, resulting in the formation of CQDs.110 Functional groups (i.e., hydroxyl, carboxyl, and amine) on the CQD surface, depending on precursor composition and reaction conditions, influence their photoluminescence properties.111 The overall process is fast, accomplished in a few minutes, and affords excellent control over the size, structure, and surface chemistry of CQDs, thus providing a very efficient and green synthesis process.103


image file: d5su00584a-f4.tif
Fig. 4 Schematic representation of the process involved in CQD formation using microwave irradiation.

Despite the extensive application potential of MAS for carbon quantum dots, current research often emphasizes the novelty of carbon sources or fluorescence performance rather than systematically addressing critical parameters governing sustainability and reproducibility. While MAS offers a rapid, energy-efficient route with excellent control over size and surface functionalities, the underlying mechanisms that link microwave parameters (power, time, and solvent polarity) with the carbonization degree and quantum yield are not consistently elucidated. Many studies focus on bio-derived or waste-based carbon sources, which are commendable from a green chemistry perspective, yet often lack comparative metrics on yield, purity, and lifecycle impacts. Moreover, although surface functionalization and heteroatom doping (e.g., N, S, and P) are reported to enhance photoluminescence and sensing capabilities, there is limited mechanistic understanding of how these modifications influence CQD behavior across applications. The diversity of precursor materials, irradiation conditions, and inconsistent reporting standards hamper scalability and reproducibility. A more structured approach, incorporating mechanistic insights, lifecycle analysis, and standardized characterization, would significantly strengthen the scientific foundation of CQD synthesis via MAS and its viability for real-world sustainable applications.

3.3 Nanocomposites

The emergence of multifunctional nanocomposites has led to a rapid development within polymer matrices owing to the improved properties they possess.158 Such composite materials are being synthesized at a fast pace in laboratories around the world for numerous applications. MAS is well known for producing nanocomposites with controlled architecture and enhanced performance.159 One such example is the development of epoxy/graphene nanocomposites using MAS (Fig. 5a). The resulting graphene layers, dispersed within the epoxy matrix, were characterized using transmission electron microscopy (TEM) and atomic force microscopy (AFM). The AFM analysis indicated that the graphene platelets were predominantly monolayer with an average thickness of 4.17 ± 0.63 nm and lateral dimensions ranging between 1.00 and 4.77 μm (Fig. 5c), confirming successful exfoliation and nanoscale uniformity.160
image file: d5su00584a-f5.tif
Fig. 5 (a) Schematic for preparation of epoxy/graphene nanocomposites, (b) schematic illustration of the synthesis of Cu3N/GDY, (c) TEM images of graphene platelets, (d) TEM and (e) AFM images of GDY, and (f) typical AFM height images of MW-rGO. Reproduced with permission from ref. 160. Copyright, 2024, RSC.

In another instance, MAS was employed to synthesize Cu3N supported on graphdiyne (GDY) (Fig. 5b), where GDY served as a two-dimensional support material. TEM and AFM imaging (Fig. 5d and e) revealed the formation of thin GDY nanosheets with a diameter of approximately 15 μm and a thickness of 4.73 nm, illustrating the structural precision achievable through microwave processing.160 These nanocomposites demonstrate enhanced surface area and nanoscale features, making them promising candidates for catalytic and energy storage applications.

Microwave-assisted synthesised polymer nanocomposites, like graphene/polypyrrole,161 PTFE/C,162 NiCoFe2O4 anchored polymer,64 hemin–graphene/poly(3,4-ethylenedioxythiophene)163 and poly(ε-caprolactone)/montmorillonite,164 have improved conductivity, mechanical strength, and antimicrobial activity, which makes them valuable for flexible electronics and biomedical applications. The following Table 2 summarizes the different nanocomposites derived via microwave irradiation and their application.

Table 2 Microwave-assisted fabrication of nanocomposites and their application
Nanocomposites Application Reference
Co3O4–graphene sheet Lithium-ion batteries 89
ZnO/MWCNT Alcohol-sensing 97
rGO–ZnO/CuO Photocatalytic degradation 165
MoS2/graphene Hydrodesulfurization 166
Graphene nanosheets–gold Electrochemical response 167
Cellulose/AgCl Antimicrobial 91
NiCoFe2O4 anchored polymer Asymmetric supercapacitor 64
Fe2O3/Ag Photocatalyst, antibacterial and anticancer 168
Cellulose/calcium silicate Biomass application 169
Graphene nanosheets–zinc sulphide Optical and photocatalytic 19
Carbon-coated Si nanocomposites Anode for lithium-ion batteries 170
Ag/ZnO/graphene Photocatalytic activity 20
Guar gum grafted sodium acrylate/cloisite Absorbent 171
N-rich carbon quantum dots/dual-phase TiO2 CO2 photoreduction 115
Pt/graphene Hydrogen peroxide sensor 172
CuO/Cu2O–ZnO mixed metal oxide Antibacterial and photocatalytic degradation 173
rGO/titania Adsorbent 174
SnO2:ZnO Photocatalytic, antimicrobial and electrochemical 175
MnCo2O4/Co3O4 Photocatalytic 176
Cellulose–silver Antimicrobial 177
Indium sulfide/indium oxide/gold Photocatalytic hydrogen generation 178
NiCo2O4–graphene oxide Supercapacitors 179
Cellulose/zinc–sulfate–calcium–phosphate (ZSCAP) Biomedical applications 180
Cobalt–polyoxometalate@carbon black Oxygen reduction reaction 158
Graphene/CoMoO4 Supercapacitors 102
Perovskite hydroxide-derived Co3O4/SnO2/reduced graphene oxide Hybrid supercapacitor devices 181
SnO2–graphite Li-ion batteries 182
Polyimides chemically cross-linked with functionalized carbon nanotubes Aerospace applications 183
BiOBr/RGO Photocatalytic activity 83
Ni@NSiC Electrocatalysis of glucose 184
AgGO Antibacterial 185
Poly(ortho-phenylenediamine-co-aniline) and functionalised carbon nanotubes Supercapacitors 186
Graphene oxide Remediation of toxic metal ions 187
CuO/MnO2 Supercapacitors 5
BiNbO4 and silver Photocatalytic 45
Carbon nitride/CuO Catalytic activity 188
Graphene–SnO2 Gas sensors 90
Mn3O4 nanoparticles@reduced graphene oxide Supercapacitors 17
BiOBr/graphene Photocatalytic activity 189
In2S3/mesoporous TiO2 Photocatalytic H2 evolution 190
Co–Fe2O3@NiO core–shell Photoelectrochemical 191
CdS/TiO2 Larvicidal activity 192
ZnO nanoparticle/poly(vinyl alcohol) Antimicrobial activity 193
Sb2O3–Ag Catalytic degradation of p-nitrophenol 101
MoS2/PANI Electrochemical 94
MWCNT/polystyrene Dielectric performance 194
Ag–ZnFe2O4@rGO Removal of organic dyes 85
NiAl-LDH/rGO Electrodes 195
Fe2O3/g-C3N4 Photocatalytic activity 196
CoFe2O4@TiO2@rGO Sonophotocatalytic degradation of tetracycline 197
PAni–SiO2 1,4-Dioxane sensor 198
MgFe2O4–ZnO Photocatalyst 87
g-C3N4/SnO2 Solar water-splitting 96
Titanium dioxide decorated graphene Photodegradation of organic dyes 86
Ag/ZnO–TiO2 Photocatalytic degradation of rhodamine B 199
Chitosan/ZnO Dye removal 6
Graphene/polypyrrole Supercapacitors 161
Gadolinium/cerium oxide High-performance supercapacitors 98
CuO/TiO2 Humidity sensor application 200
TiO2/graphene Photocatalytic activity for methylene blue degradation 201
Silver nanoparticles and chitosan Removal of reactive blue-19 202


The rapid advancement of MAS in fabricating multifunctional nanocomposites underscores its role as a powerful tool in sustainable material design. However, much of the current literature tends to present MAS-derived composites as universally superior without fully interrogating the scalability, reproducibility, or long-term stability of these systems. While MAS can yield nanocomposites with enhanced electronic, mechanical, or catalytic properties, there is often limited discussion on how these improvements quantitatively compare to composites synthesized via conventional routes. Furthermore, the diverse range of materials (e.g., polymers, metal oxides, and carbon-based supports) and hybrid structures reported, though impressive, lacks standardization in synthesis conditions and performance benchmarking, making it difficult to assess their broader applicability. The role of interfacial interactions, dispersion uniformity, and matrix compatibility in determining composite performance is also underexplored, especially when natural or bio-derived materials are used. A more critical and comparative analysis of reaction parameters, material interfaces, and environmental impact, particularly for composites intended for biomedical, energy, or environmental applications, is essential for establishing MAS as a truly sustainable nanocomposite fabrication strategy.

4 Comparative analysis with other sustainable methods

To critically evaluate the position of MAS within the broader landscape of sustainable synthesis, it is essential to compare its performance with other established sustainable methods including sonication, electrochemical synthesis, and enzymatic approaches. Table 3 summarizes a comparative analysis of different sustainable methods available for the synthesis of nanomaterials. MAS operates through dielectric heating mechanisms that provide rapid and uniform heating, typically achieving superior reaction rates with most reactions completing within minutes to hours.12 In contrast, sonochemical methods rely on acoustic cavitation to create localized high-energy zones, often requiring longer time frames but excel in operational simplicity and cost-effectiveness.203 While MAS demonstrates significant advantages through its selective heating mechanism and broad substrate compatibility, sonication offers better scalability for industrial applications through continuous flow systems that can be more readily implemented than large-scale microwave reactors. Electrochemical synthesis presents a fundamentally different sustainable approach through electron transfer mechanisms, offering precise control over reaction conditions that MAS cannot match.204 While MAS provides rapid heating and acceleration of thermal processes, electrochemical methods enable selective redox transformations under mild conditions without chemical oxidants or reductants, achieving high atom economy and eliminating chemical waste. However, electrochemical synthesis faces limitations in substrate scope, being restricted to compounds with suitable electrochemical properties, whereas MAS demonstrates broader applicability across various chemical transformations. The scalability of electrochemical methods often surpasses MAS for industrial applications, though MAS maintains advantages in reaction versatility and processing speed across diverse chemical classes.
Table 3 Summary of different sustainable methods applicable for nanomaterial fabrication
Parameter Microwave-assisted Sonication/ultrasonic Electrochemical Enzymatic
Activation mechanism Dielectric heating through electromagnetic radiation Acoustic cavitation through ultrasonic waves (>20 kHz) Electron transfer through electrical current Biocatalytic conversion through enzymes
Reaction time Significantly reduced (minutes to hours) Shortened compared to conventional methods Variable, depends on the current density Generally longer but mild conditions
Energy efficiency High efficiency with selective heating Moderate to high efficiency High efficiency, electricity-driven Very high efficiency under mild conditions
Reaction temperature Mild to moderate temperatures Ambient to moderate temperatures Ambient to moderate temperatures Ambient to mild temperatures (30–80 °C)
Selectivity High selectivity and yields Enhanced selectivity through cavitation High selectivity with controlled potential Excellent selectivity and stereoselectivity
Environmental impact Green method, reduced solvent use Environmentally friendly, minimal waste Green alternative when using renewable electricity Highly sustainable, biodegradable catalysts
Scale-up potential Moderate to good scalability Good scalability Excellent scalability Limited scalability, cost considerations
Equipment cost Moderate initial investment Low to moderate cost Moderate to high cost Low equipment cost, high enzyme cost
Solvent requirements Reduced solvent use, water-compatible Compatible with various solvents Often aqueous or low-toxicity solvents Aqueous systems preferred
Product purity High purity, fewer side products Good purity with controlled conditions High purity with selective reactions Excellent purity, minimal purification
Substrate scope Wide range of substrates Broad substrate compatibility Limited by electrochemical properties Limited by enzyme specificity
Operational simplicity Simple operation Simple operation Moderate complexity Moderate complexity


Enzymatic synthesis represents the pinnacle of selectivity in sustainable chemistry, operating under the mildest conditions with exceptional stereoselectivity and minimal side reactions using biodegradable, renewable catalysts.205 While MAS achieves rapid reaction rates through thermal activation, enzymatic methods accomplish transformations at ambient temperatures with precision that MAS may struggle to achieve under rapid heating conditions. The sustainability comparison heavily favors enzymatic methods due to their mild aqueous conditions and renewable nature,206 though MAS demonstrates significant advantages in reaction speed and substrate versatility. The cost analysis reveals complexity: MAS requires moderate initial equipment investment, while enzymatic methods face ongoing costs related to enzyme procurement and replacement, with substrate scope limitations restricted by specific enzyme–substrate interactions, contrasting sharply with MAS's broad applicability.

The comparative analysis reveals that no single sustainable method dominates across all parameters, with MAS excelling in rapid processing and broad substrate compatibility207 while facing limitations in uniform heating for large-scale reactions and requirements for microwave-transparent materials. The integration of multiple sustainable methods presents promising opportunities, with hybrid approaches combining MAS with sonication for enhanced mixing or with enzymatic steps for improved selectivity, potentially addressing individual method limitations. The critical evaluation indicates that MAS represents a significant advancement in green chemistry, but its position within the sustainable synthesis landscape is complementary rather than dominant, with true value emerging when applied appropriately to reactions where rapid heating and broad substrate compatibility are priorities, while recognizing situations where alternative sustainable methods may provide superior performance.

5 Green metrics and sustainability assessment

In alignment with the green chemistry principles, a set of rules meant to support environmentally friendly and financially viable chemical operations, MAS provides a cleaner, more sustainable pathway for the synthesis of nanomaterials by minimizing waste, lowering reaction temperatures, and enhancing atom economy.13,15,208 Prevention of waste: employing improved control over reaction parameters in MAS minimizes the generation of undesired by-products, thus lowering waste at the source. Atom economy: a crucial component of atom-economical processes, MAS frequently encourages high-yield reactions with fewer steps, optimizing the incorporation of all starting materials into the finished product.207 Energy efficiency: MAS usually requires lower temperatures and shorter durations, resulting in significant energy savings, because of the quick internal heating made possible by microwaves.34 Use of safer solvents and auxiliaries: to promote safer laboratory and industrial procedures, many MAS protocols make use of safe solvents like ethanol or water, and in certain situations, they completely do away with the need for solvents.52 Design for degradation: MAS makes it easier to synthesize materials that are more environmentally friendly throughout their life cycle when combined with biodegradable precursors like plant extracts or renewable feedstocks.145 MAS is a sustainable chemical manufacturing model in addition to being an effective synthetic tool. It has the potential to address urgent global issues such as energy consumption, pollution, and sustainable development in the field of nanomaterials because it is in line with important green chemistry principles.

To evaluate the environmental and process efficiency of MAS methods, green metrics serve as critical tools.209,210 Numerous quantitative green metrics have been developed to objectively evaluate MAS's environmental sustainability. These metrics are used to assess how well a chemical process conforms to the green chemistry principles, especially concerning energy consumption, waste generation, and resource efficiency.211 The benefits of MAS over traditional synthesis methods are demonstrated by evaluating it using these standardized parameters.

Fig. 6 illustrates a schematic overview of the key green metrics, highlighting their role in quantifying the eco-friendliness of MAS processes. These metrics help in comparing MAS with conventional synthetic strategies and support the optimization of greener routes. Table 4 summarizes these green metrics, providing definitions and the specific advantages offered by MAS over conventional methods. It demonstrates how MAS routes often achieve lower waste generation, higher material utilization, and superior energy performance, thereby supporting sustainable development in nanomaterials synthesis.


image file: d5su00584a-f6.tif
Fig. 6 Green metrics involved in MAS.
Table 4 Summary of green metrics and their relevance in MAS
Green metric Relevance in MAS Example/observation
E-factor (environmental factor) MAS minimizes waste due to efficient conversions and minimal use of solvents Carbon nitride-supported silver nanoparticles: E-factor is 0.27 (ref. 211)
Atom economy (%) Promotes one-pot reactions and reduces the formation of by-products MAS often results in higher atom economy due to fewer side reactions. Example: CQDs from fenugreek seeds127
Process mass intensity (PMI) Lower PMI values reflect efficient use of resources and reduced waste MAS offers lower PMI due to shorter reaction and minimal solvent use
Energy efficiency MAS significantly reduces energy consumption due to volumetric heating and shorter reaction times MAS of TiO2 showed ∼40% lower energy consumption vs. the sol–gel method
Reaction time MAS reduces reaction time from hours to minutes in most cases Metal nanoparticle synthesis is often completed within 5–30 minutes under MAS15
Solvent use MAS supports green solvents like water and ethanol, or even solvent-free synthesis Silver and gold nanocatalysts from Aerva lanata extract under MAS with water as the solvent55
Temperature profile MAS allows low to moderate temperature reactions with precise control, avoiding thermal degradation Hylocereus costaricensis-mediated Ag/ZnO-NPs synthesis was perfromed under mild MAS conditions75


5.1 E-factor (environmental factor)

The E-factor is a widely accepted metric used to quantify the amount of waste generated per unit of product.210 It is calculated using the formula:
image file: d5su00584a-t1.tif

The E-factor in MAS is substantially lower than that in traditional techniques.211 This is mostly because fewer solvents are used, fewer side products are produced, and the reaction pathways are shorter.212 For example, the microwave approach showed an E-factor reduction of more than 60% in comparison to conventional thermal methods in a comparative study involving the synthesis of ZnO nanoparticles.213 A smaller environmental impact and streamlined post-synthesis purification procedures are the direct results of this reduction.

5.2 Atom economy (%)

Atom economy is a concept describing how much of the reactants are used for the final product and is given by,210
image file: d5su00584a-t2.tif

MAS usually promotes high atom economy as most reactions are conducted in one-pot and solvent-free, minimizing excess reagent and the generation of unwanted by-products.212 Thus, a higher ratio of starting materials becomes part of the end nanomaterials, both improving sustainability and economic efficiency in the process.

5.3 Process mass intensity (PMI)

Process mass intensity (PMI) is another valuable metric used to determine the material efficacy of a synthesis route and is given by the relation,210
image file: d5su00584a-t3.tif

MAS typically produces a lower PMI than traditional methods. This is attributable to the low solvent volume, short reaction time, and efficient heating methods that enable quicker and more complete reactions.34 An effective PMI value demonstrates a tight process with minimal waste of materials, which can help develop nanomaterials more sustainably on a large scale.

5.4 Energy efficiency

The most noteworthy benefit of MAS is its improved energy efficiency. Microwave systems enable the fast and specific heating of reaction mixtures, minimizing the energy losses characteristic of bulk heating systems.23 Interestingly, MAS reportedly consumes 80% less energy compared to conventional hydrothermal synthesis.40 This is mainly because the energy delivered is directly coupled to the reactants, as opposed to heating the entirety of the vessel or the surrounding medium. Additionally, one major factor that keeps MAS as a pioneering energy-efficient synthesis method is the accurate control of microwave power, preventing overheating and wastage of energy.22

5.5 Solvent and reagent aspects

One of the main advantages of MAS is that it can easily be used with green solvents and natural reagents, which means that MAS is itself an inherently sustainable method.212 Compared with traditional synthesis processes that typically require toxic solvents or harmful surfactants, MAS allows the use of eco-friendly and renewable materials without loss of process efficiency or product quality.46,128,169 Water is likely the most common solvent amongst all solvents used for MAS because of its high dielectric constant, which enables it to interact strongly with microwave radiation.214 This allows the microwave energy to be absorbed very effectively, allowing for very rapid and uniform heating. In addition, water is non-toxic, non-flammable, easily obtainable, and compatible with many biological precursors, making it an ideal medium for eco-friendly synthesis.215,216 MAS in aqueous media has proven to be an effective strategy for producing nanocrystalline three-dimensional metal oxides and gold nanoparticles.217 This approach not only enhances the structural and morphological control of the materials but also improves their functional performance in various applications.

Organic salts known as ionic liquids have low melting points and exhibit novel solvent and thermal stability properties.218 One area where MAS using ionic liquids as reaction media has several advantages is that it improves the solubility of reactants, improves mass transfer and reduces energy consumption.219 Additionally, ionic liquids can act as a solvent and stabilizing agent, and no additional toxic solvents or surfactants are necessary. Nanoparticles such as MgO,46 ZnO,220 ZnS quantum dots221 and carbon dots222 have been prepared via this approach in the presence of the respective ionic liquids, showing enhanced applications in various fields. The green nanomaterial synthesis using MAS in ionic liquids has been an attractive approach because ionic liquids can produce nanomaterials with tailored properties and high purity.218,219

Another frequently employed green solvent in MAS is ethanol, which has moderate polarity and a relatively high boiling point.223 Being renewable and biodegradable, ethanol from biomass is consistent with the green chemistry principles. In particular, the solvating power of the ionic liquids toward various kinds of metal precursors and the capability to work with various plant-derived extracts make them suitable for the nanomaterial or nanocomposite synthesis.217

An exciting advance in the field of sustainable nanomaterial synthesis, including MAS is using plant extracts, which act as reducing agents, stabilizers, and capping agents.224 The extracts are bioactive in nature and contain species like flavonoids, terpenoids, tannins, polyphenols and alkaloids.75 They serve multiple functions, so there is no need to use toxic reagents and synthetic surfactants that are used to manipulate nanoparticle morphology and stability. Apart from environmental safety, the incorporation of plant extracts as reagents in MAS often endows the synthesized nanomaterials with additional functional properties. Examples include increasing the antioxidant activity of the nanoparticle surface by using phytochemical residues, improving biological interactions and enhancing their dispersion stability in aqueous systems.41 This paves the way for applications in biomedicine, food packaging and environmental remediation. The thoughtful selection of solvents and reagents in MAS contributes significantly to making the process more sustainable. By incorporating renewable, non-toxic, and multifunctional materials, MAS not only reduces the environmental impact but also improves the performance and applicability of the resulting nanomaterials.

5.6 Lifecycle assessment (LCA) and environmental impact

Lifecycle assessment (LCA) is an analytical approach used to determine the environmental impacts of a process, product, or technology from beginning to end-including raw material sourcing, production, usage, and end-of-life treatment.225 In MAS, LCA is valuable for measuring factors such as energy usage, emissions, waste output, and overall resource efficiency.226 The assessment typically involves four key stages:226 (i) defining the goal and scope, which sets the purpose and boundaries of the analysis; (ii) inventory analysis, where input and output data (like materials, energy, and waste) are gathered; (iii) impact assessment, in which environmental consequences such as the carbon footprint or toxicity are evaluated; and (iv) interpretation, where findings are analyzed to identify sustainable improvements and guide greener decision-making. Software tools such as SimaPro,225 GaBi,227 and OpenLCA,228,229 allow for detailed modelling, data analysis, and impact assessment based on international standards. These tools help ensure standardized, transparent, and repeatable environmental assessments aligned with ISO 14040 and ISO 14044 guidelines.230 These tools enable users to evaluate the environmental impacts of products or processes throughout their life cycle.

LCA of ZnO nanostructures synthesized using a MAS technique with varying microwave power (110 W to 710 W) and annealing temperature (90 °C to 220 °C) revealed that increasing microwave power or synthesis temperature led to enhanced material output with lower environmental impact, and the footprint reduced by 27% (power) and 41% (temperature).231 Silver nanoparticles (AgNPs), synthesized using MAS with glucose as the reducing agent and food-grade corn starch as the stabilizer, were evaluated using LCA in GaBi 6.0 software.227 The analysis revealed that the primary environmental burdens arose from silver metal production and electricity consumption. In contrast, glucose and starch contributed negligible impacts, affirming the environmental benefits of using biogenic, food-safe reagents.

A LCA study compared the environmental impacts of six bottom-up synthesis methods for carbon dots (CDs), focusing on hydrothermal and MAS using citric acid and urea.232 The results revealed that electricity use dominated the impacts in hydrothermal methods, while citric acid was the major contributor in MAS. When adjusted for quantum yield, MAS using both citric acid and urea emerged as a more sustainable option for high-performance CD production.232 Fernandes et al. conducted a LCA study comparing high-yield and standard bottom-up methods for fabricating carbon dots, using MAS.233 Their findings revealed that certain standard methods, notably MAS, exhibited lower environmental impacts compared to some high-yield conventional techniques. These studies suggest that MAS offers a more sustainable approach for carbon dot production.

Nanocomposites of cerium oxide, copper oxide, and zinc oxide (CeO2–CuO–ZnO; CeCZ) were developed through MAS across a range of 300 to 800 W.234 An LCA based on power consumption indicated that utilizing lower microwave energy notably decreases greenhouse gas emissions and lessens the environmental footprint. The findings further suggest that incorporating renewable energy into the process can enhance its sustainability. This study demonstrates the effectiveness of MAS as a green and energy-saving technique for creating advanced photocatalysts aimed at environmental cleanup. A review of J. Munuera et al. explores multiple sustainable approaches for graphene synthesis, with a particular focus on MAS due to its energy efficiency and lower environmental burden compared to conventional methods.228 It underscores the role of LCA in evaluating the ecological impacts of these production techniques, offering a comprehensive view of their sustainability.

LCA indicates significant environmental benefits of MAS over conventional synthesis methods. Moreover, MAS processes are often solvent-free or use minimal solvents, resulting in reduced wastewater generation and a lower burden of effluent treatment. Furthermore, in closed-vessel microwave systems, chemicals are contained, and the risk to the user from chemical exposure and release of volatile organic compounds (VOCs) is reduced.218 Thus, MAS is efficient, clean, and meets the regulatory mandates for safe and environmentally friendly manufacturing processes.

5.7 Circular chemistry and resource efficiency

MAS aligns well with the tenets of circular chemistry, a framework that seeks to limit the use of resources, waste, and use waste as feedstock for useful products.209 MAS advocates for this model in multiple ways. MAS can efficiently utilize plant-derived precursors, such as fruit peels,81 leaf extracts,26 and agricultural waste.235 These biomaterials, which are usually high in reducing agents and stabilizing agents, are good substitutes for toxic chemical reagents. MAS enables the transformation of biomass waste into valuable nanomaterials.18,121,236 Various metal and metal oxide nanoparticles particularly silver,237 gold,238 zinc oxide,75 and copper oxide,239 have been synthesized using plant extracts of Aloe vera, Moringa oleifera, and Azadirachta indica under microwave irradiation, and exhibit astounding antibacterial, catalytic, and optical activities. Some other materials, which can potentially become agro-wastes, are not only collected but also converted into functional materials, e.g., carbon quantum dots (CQDs) and doped metal oxides have been successfully synthesized from sugarcane bagasse,113 citrus peels,108 and other biogenic residues.

The reduced use of organic solvents and progress of reactions contained in sealed vessels lowers emissions and avoids cross-contamination, thus contributing to the novel trend of circular economy, and green engineering principles.15,214 The industrial capacity of MAS is becoming more and more understood, particularly due to its capacity to manufacture high-quality nanomaterials at a low environmental cost. Case studies show the scalability of MAS and its implementation in sustainable manufacturing. Carbon nanospheres, synthesised using banana peel extract under microwave conditions, were applied for selective and specific sensing of Cu2+ ions and the amino acid tryptophan in aqueous solutions.240 The use of agricultural waste as a precursor and a capping agent eliminated toxic chemicals. Photoluminescent carbon dots obtained from palm frond biomass wastes via MAS was demonstrated to be an exceptional sensor for lead(II) ions.236 Such examples reinforce the relevance of MAS not only in laboratory-scale research but also in industrial applications where eco-efficiency, performance, and cost-effectiveness are equally critical. The adoption of MAS supports the development of nanotechnologies that are not just advanced in function, but also responsible in practice-fostering a future where innovation and environmental stewardship go hand in hand.

6 Applications of microwave-assisted nanomaterials

Due to their unique properties at the nanoscale, nanomaterials have gained considerable interest in multiple fields due to their potential to transform industries and address key societal challenges. The advantages of MAS include rapid synthesis kinetics, scalability and ability to produce nanomaterials with high purity and uniformity for electronics, catalysis and biomedical engineering applications (Fig. 7). These nanomaterials are proved to be versatile solutions, possessing improved performance and new functionalities for biomedical applications, environmental remediation and energy storage.26 Moreover, they are important in the drive towards electronics, optoelectronics, industrial applications of miniaturization, better power efficiency and multifunctionality.
image file: d5su00584a-f7.tif
Fig. 7 Applications of microwave-assisted fabricated nanomaterials.

6.1 Biomedical applications

Because of their unique and versatile properties, nanomaterials have revolutionized biomedical research and healthcare.2 Some key applications include drug delivery, imaging, therapeutics and biosensors. Encapsulating drugs, peptides or nucleic acids in nanoparticles enables the targeting of these to specific tissues or cells.79 Controlled release and enhanced therapeutic efficacy are realized by surface modification to minimize side effects.241 As contrast agents for a variety of imaging modalities, including magnetic resonance imaging (MRI),242 bio-imaging,136 cell-imaging107 or fluorescence imaging,114,121 nanomaterials, such as quantum dots, iron oxide nanoparticles or gold nanoparticles, are used. The cellular and chemical properties of nanomaterials are being exploited to design nanosystems for use in therapeutic drug monitoring,142 antibacterial therapy,157 photodynamic therapy,27 cancer therapy,144 infections, and other diseases.79 Targeted delivery systems increase treatment efficacy and decrease systemic toxicity. Biosensors based on nanomaterials provide fast and sensitive biomolecule, pathogen, and disease marker detection for diagnosis.243 Nanoparticles are functionalized with disease targets for signal transduction acceleration and sensor enhancement.135,172,198

Microwave-assisted synthesis has accelerated the development of nanomaterials tailored for biomedical applications, offering rapid, energy-efficient fabrication of highly uniform and functionalized nanoparticles. However, while MAS-generated nanomaterials have shown promise in drug delivery, imaging, therapeutics, and biosensing,12 much of the existing literature tends to emphasize performance outcomes without adequately addressing long-term biocompatibility, in vivo stability, or toxicological implications. For instance, although quantum dots and metal-based nanoparticles synthesized via MAS exhibit superior imaging and therapeutic properties, concerns about their degradation, bioaccumulation, and clearance from biological systems remain largely underexplored. Moreover, the surface functionalization of nanoparticles, while improving targeting and controlled release, is often not correlated with reproducibility or consistency across batches, an essential factor for clinical translation. Biosensors based on MAS nanomaterials have demonstrated heightened sensitivity and specificity, yet few studies address their real-world diagnostic reliability under complex biological conditions. There is also a gap in comparing MAS-derived nanomaterials with those synthesized through traditional methods in terms of cost-effectiveness, lifecycle analysis, and regulatory compliance for medical use. Therefore, a critical focus on bridging the gap between laboratory performance and clinical feasibility, backed by standardized testing protocols, is necessary to validate MAS as a sustainable and scalable route for biomedical nanomaterial development.

6.2 Environmental remediation

Utilizing nanomaterials makes for innovative solutions to environmental remediation and pollution control through pollutant removal, detection and monitoring.2 Key applications include water purification, air filtration and environmental sensing.126,145,238 For the removal of heavy metals,113,244 organic pollutants,80,107 and pathogens from water sources, nanomaterials, like graphene oxide, carbon nanotubes, and metal oxide nanoparticles, are employed in adsorption, filtration and photocatalysis.1 Particulate matter, volatile organic compounds, and harmful gases are captured from indoor and outdoor air by nanomaterial-based filters and membranes that improve indoor and outdoor air quality and decrease health risks. Real-time sensors and detectors based on nanomaterials make it possible to monitor pollutants, greenhouse gases, and environmental contaminants in air, water and soil for early warning systems and to develop environmental management strategies.145,148

Microwave-assisted synthesis of nanomaterials for environmental remediation holds immense promise due to the ability to rapidly produce high-surface-area materials with tailored physicochemical properties for pollutant capture and degradation. However, despite the growing number of studies highlighting their efficacy in water purification, air filtration, and environmental sensing, many claims remain largely performance-based, lacking comprehensive evaluation of environmental safety, reusability, and post-treatment disposal. While materials like graphene oxide, carbon nanotubes, and metal oxides synthesized via MAS demonstrate efficient removal of heavy metals and organic pollutants, there is insufficient focus on their regeneration potential and cost-effectiveness in real-world applications. Future work should critically address the long-term sustainability, risk assessment, and policy implications of deploying MAS-derived nanomaterials in environmental technologies.

6.3 Energy applications

Renewable energy technologies require and benefit from the use of nanomaterials to advance their impact. Major applications include solar cells, energy storage, and catalysis.12 Next generation solar cells use nanomaterials like quantum dots, perovskites, and nanowires, which increase light absorption, charge transport, and device performance capabilities of solar cells.3,236,245,246 Also, nanostructured materials permit the fabrication of lightweight, flexible, and low-cost photovoltaic devices.14 To improve energy density, cycling stability and charge–discharge rates, nanomaterials, including carbon nanotubes, graphene and metal oxides are used in lithium ion batteries, supercapacitors, and fuel cells.12,247 This leads to energy storage and electrochemical performance enhancement of the electrodes and electrolytes via nanostructuring.64,181 Because of their efficiency, nanomaterials are indeed excellent catalysts to conduct chemical reactions in energy conversion and energy storage processes, including hydrogen production, synthesis of fuels, and pollutant degradation.14,82,101,248 Enriched nanostructures and active sites enhance the catalytic activity and selectivity.

While MAS-derived nanomaterials show promise in improving light absorption and energy storage performance, real-world testing under variable conditions and lifecycle assessments are rarely addressed. Furthermore, the long-term stability and degradation behavior of these materials in operational environments remain critical gaps that hinder their commercialization. A more application-oriented and techno-economic evaluation is needed to validate MAS as a truly transformative tool in the renewable energy sector.

6.4 Electronics and optoelectronics

Modern electronic devices and optoelectronic technologies are dependent on and require nanomaterials in order to facilitate miniaturization, high performance, and new functionalities. Major applications include nanoelectronics, flexible electronics, and photonics.3 To achieve high speed operation, low power consumption and improved electronic properties, MAS based nanomaterials, such as carbon nanotubes, graphene and semiconductor nanoparticles, are used in transistors, supercapacitors, sensors and memory devices.2 The development of flexible displays, wearable sensors, and electronic textiles is enabled by nanomaterial-based flexible substrates, conductive inks and transparent electrodes.98,200 As lightweight, bendable and stretchable electronic devices, nanocomposites and thin-film technologies are being developed. Optical devices based on nanomaterials, such as plasmonic nanoparticles, quantum dots, and photonic crystals, are used in photonic circuits, LEDs, and display technology with applications in telecommunications and sensing.4,27,103 With nanostructured materials, precise control of light–matter interactions and optical properties is achieved (Fig. 8).249
image file: d5su00584a-f8.tif
Fig. 8 Schematic representation of MAS of MnSeTe nanocomposite flowers for optoelectronic and photoresponse applications. Reproduced with permission from ref. 249. Copyright, 2024, ACS.

However, despite promising laboratory outcomes in flexible electronics, memory devices, and photonic applications, the practical integration of MAS-derived nanomaterials into large-scale device fabrication remains a major hurdle. Challenges such as batch-to-batch consistency, substrate compatibility, and thermal stability under operational conditions are often overlooked. Moreover, while nanostructured materials enable improved optical and electronic properties, their long-term reliability, particularly in flexible and wearable formats, is not rigorously evaluated. Future research must bridge the gap between material innovation and real-world device engineering by focusing on durability, manufacturability, and cost-benefit analyses.

6.5 Industrial applications

Nanomaterials are used in many industrial sectors to improve material performance, and enable new product properties and functionalities. Major applications include advanced materials, textiles, packaging, food and agriculture.2 Microwave-based nanomaterials that include nanocomposites, nanocoatings and nanoporous materials are used to strengthen mechanical strength, thermal stability and chemical resistance for automotive, aerospace and construction industries.3 Lightweight, durable and multifunctional materials for structural components and coatings can be developed using nanostructured materials. Textiles, clothing, and packaging materials use coatings, films and fibers containing nanomaterials to resist stains, be antimicrobial (Fig. 9)250 and to serve as barrier materials for moisture, gases and UV radiation.251 Nanostructured materials increase durability, comfort, and functionality of textiles and packaging products.88,252 In food packaging, processing, and storage, MAS based nanomaterials are employed to enhance the safety, preservation, and quality of food.2,35,147 Controlled release of antimicrobial agents, antioxidants and nutrients by nanoemulsions, nanocapsules, and nanosensors reduces food spoilage and contamination. In many instances, the applications in these materials represent just a fraction of the potential of nanomaterials to address societal challenges, where transformative technologies are critical to progress and to improve quality of life across numerous sectors.
image file: d5su00584a-f9.tif
Fig. 9 Proposed mechanism of the cotton fabrics treated with the Ag/ZnO nanocomposite for antibacterial activity. Reproduced with permission from ref. 250. Copyright, 2022, RSC.

While applications in textiles, packaging, and construction have leveraged MAS-derived nanocomposites and coatings for added durability, antimicrobial activity, and environmental resistance, many studies lack long-term performance evaluations under real industrial conditions. Moreover, the regulatory and safety implications of incorporating nanomaterials, particularly in food packaging and consumer goods, are often underreported or insufficiently addressed. The controlled release of bioactive agents in food systems, while conceptually valuable, still faces hurdles related to migration, degradation, and human health impacts. For MAS to truly support transformative industrial innovation, future research must critically integrate lifecycle assessment, regulatory compliance, and real-world performance data to validate both safety and sustainability claims.

7 Challenges and future prospects

Despite these advantages, MAS still faces several critical limitations that hinder its widespread adoption, particularly at the industrial scale. One major challenge is the difficulty in scaling up the process without compromising product uniformity or energy efficiency. In laboratory-scale setups, the small reaction volumes ensure uniform microwave absorption, but scaling to larger volumes often results in uneven heating and local hot spots, which can lead to variability in particle size, shape, and crystallinity. Furthermore, many published studies report successful synthesis under highly controlled lab conditions, but often lack discussions on reproducibility, batch-to-batch consistency, or adaptability to continuous production models. This disconnect between academic success and industrial feasibility must be addressed through engineering innovations in reactor design and process standardization.

Another limitation is the incomplete mechanistic understanding of microwave–material interactions. Unlike conventional heating, where temperature and reaction pathways are more predictable, MAS involves complex electromagnetic interactions that are not yet fully understood. This knowledge gap hinders the rational design of synthesis protocols and limits control over fine-tuned nanostructures. Moreover, while natural and bio-based precursors are widely used, their compositional variability, due to seasonality, geography, and storage conditions, poses a challenge for reproducibility. Therefore, future research should emphasize reaction mechanism elucidation, standardization of green feedstocks, and the development of robust synthesis protocols that can adapt to feedstock variability.

Looking ahead, the future of MAS depends on strategic, interdisciplinary efforts. The development of scalable, energy-efficient, and automated microwave reactors tailored for nanomaterial synthesis is crucial. Integration of artificial intelligence (AI) and machine learning (ML) tools can revolutionize MAS by enabling predictive modelling, real-time process control, and adaptive optimization based on feedback from reaction parameters. Additionally, the environmental claims of MAS need to be backed by rigorous life cycle assessments (LCA) and toxicological evaluations to ensure long-term sustainability and safety. Cross-sector collaboration between chemists, biologists, materials scientists, engineers, environmental experts, and industry leaders is essential for transitioning MAS from a promising laboratory technique to a widely adopted industrial technology.

8 Conclusion

Microwave-assisted synthesis (MAS) stands out as a breakthrough approach in the field of sustainable nanotechnology, offering a rapid, energy-conscious, and eco-friendly method for creating nanomaterials with controlled properties. This review has explored the core principles that govern MAS, its strong alignment with green chemistry, and its broad utility in the synthesis of metallic nanoparticles, carbon quantum dots (CQDs), and nanocomposites. The distinctive mode of microwave heating-characterized by uniform, internal, and selective energy delivery-enables faster reactions, improved yields, and reduced processing times under gentler conditions when compared to conventional techniques. A particularly noteworthy feature of MAS is its compatibility with green synthesis approaches. Utilizing natural and renewable precursors such as plant extracts, biodegradable materials, and bio-based reagents not only decreases the environmental footprint but can also impart valuable functional traits to the synthesized nanomaterials such as improved biocompatibility, catalytic activity, or targeted responsiveness. The materials produced through MAS have demonstrated great potential across various fields, including catalysis, drug delivery, environmental remediation, biosensing, and renewable energy systems. Considering its sustainable and eco-friendly properties, MAS has a considerable advantage over other sustainable methods and conventional methods, and is capable of driving the future of sustainable nanomaterial manufacturing.

Author contributions

Sajini T.: conceptualization, writing – original draft, supervision. Jebin Joseph: literature review, writing – review and editing.

Conflicts of interest

The authors declare no conflict of interest in this paper.

Data availability

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

References

  1. S. Rathod, S. Preetam, C. Pandey and S. P. Bera, Biotechnol. Rep., 2024, 41, e00830 CrossRef PubMed.
  2. A. I. Osman, Y. Zhang, M. Farghali, A. K. Rashwan, A. S. Eltaweil, E. M. Abd El-Monaem, I. M. A. Mohamed, M. M. Badr, I. Ihara, D. W. Rooney and P. S. Yap, Environ. Chem. Lett., 2024, 22, 841–887 CrossRef.
  3. O. Gohar, M. Zubair Khan, I. Bibi, N. Bashir, U. Tariq, M. Bakhtiar, M. Ramzan Abdul Karim, F. Ali, M. Bilal Hanif and M. Motola, Mater. Des., 2024, 241, 112930 CrossRef CAS.
  4. A. Sakr, A. Kamal, L. Jacob, R. Homsi and H. Butt, Polym. Rev., 2025, 65, 1–46 CrossRef.
  5. P. Zhang and W. Li, Micro Nano Lett., 2020, 15, 938–942 CrossRef.
  6. M. H. Mostafa, M. A. Elsawy, M. S. A. Darwish, L. I. Hussein and A. H. Abdaleem, Mater. Chem. Phys., 2020, 248, 122914 CrossRef CAS.
  7. S. Supriya, S. Das, S. K. Samal, S. Senapati and R. Naik, Nanoscale, 2024, 16, 7566–7581 RSC.
  8. M. A. B. Turki and T. A. Salman, J. Chem. Rev., 2024, 6, 458–484 CAS.
  9. S. Dhahri, H. Shall and N. Thabet Mliki, Inorg. Chem. Commun., 2024, 168, 112872 CrossRef CAS.
  10. M. A. Dheyab, N. Oladzadabbasabadi, A. A. Aziz, P. M. Khaniabadi, M. T. S. Al-ouqaili, M. S. Jameel, F. S. Braim, B. Mehrdel and M. Ghasemlou, J. Environ. Chem. Eng., 2024, 12, 112345 CrossRef.
  11. A. K. Nagariya, A. K. Meena, A. K. Yadav, U. S. Niranjan, A. K. Pathak, B. Singh and M. M. Rao, J. Pharm. Res., 2010, 3, 575–580 Search PubMed.
  12. N. Devi, S. Sahoo, R. Kumar and R. K. Singh, Nanoscale, 2021, 13, 11679–11711 RSC.
  13. J. Wojnarowicz, T. Chudoba and W. Lojkowski, Nanomaterials, 2020, 10, 1086 CrossRef CAS PubMed.
  14. T. V. De Medeiros, J. Manioudakis, F. Noun, J. R. Macairan, F. Victoria and R. Naccache, J. Mater. Chem. C, 2019, 7, 7175–7195 RSC.
  15. A. Jabbar Khan, L. Gao, A. Numan, S. Khan, I. Hussain, M. Sajjad, S. S. Shah, A. Mateen and G. Zhao, Crit. Rev. Solid State Mater. Sci., 2025, 1–24 Search PubMed.
  16. K. Kannan, D. Radhika, M. P. Nikolova, V. Andal, K. K. Sadasivuni and L. S. Krishna, Optik, 2020, 218, 165112 CrossRef.
  17. X. She, X. Zhang, J. Liu, L. Li, X. Yu, Z. Huang and S. Shang, Mater. Res. Bull., 2015, 70, 945–950 CrossRef.
  18. M. Nazar, M. Hasan, B. Wirjosentono, B. A. Gani and C. E. Nada, ACS Omega, 2024, 9, 20571–20581 CrossRef PubMed.
  19. H. Hu, X. Wang, F. Liu, J. Wang and C. Xu, Synth. Met., 2011, 161, 404–410 Search PubMed.
  20. P. Dou, F. Tan, W. Wang, A. Sarreshteh, X. Qiao, X. Qiu and J. Chen, J. Photochem. Photobiol., A, 2015, 302, 17–22 CrossRef.
  21. V. Polshettiwar, M. N. Nadagouda and R. S. Varma, Aust. J. Chem., 2009, 62, 16–26 CrossRef.
  22. C. Valverde, M. M. Rodríguez-García, E. Rojas and R. Bayón, Int. Commun. Heat Mass Transf., 2024, 156, 107594 CrossRef.
  23. E. T. Thostenson and T. W. Chou, Composites, Part A, 1999, 30, 1055–1071 Search PubMed.
  24. H. Li, Z. Zhao, C. Xiouras, G. D. Stefanidis, X. Li and X. Gao, Renewable Sustainable Energy Rev., 2019, 114, 109316 CrossRef CAS.
  25. J. Sun, W. Wang and Q. Yue, Materials, 2016, 9, 231 CrossRef PubMed.
  26. S. Yalshetti, B. Thokchom, S. M. Bhavi, S. R. Singh, S. R. Patil, B. P. Harini, M. Sillanpää, J. G. Manjunatha, B. S. Srinath and R. B. Yarajarla, Sci. Rep., 2024, 14, 1–13 Search PubMed.
  27. M. P. Romero, F. Alves, M. D. Stringasci, H. H. Buzzá, H. Ciol, N. M. Inada and V. S. Bagnato, Front. Microbiol., 2021, 12, 1–13 Search PubMed.
  28. A. V. Murugan, T. Muraliganth and A. Manthiram, J. Phys. Chem. C, 2008, 112, 14665–14671 Search PubMed.
  29. N. Moghni, H. Boutoumi, H. Khalaf, N. Makaoui and G. Colón, J. Photochem. Photobiol., A, 2022, 428, 113848 CrossRef CAS.
  30. R. Rosa, L. Trombi, P. Veronesi and C. Leonelli, Int. J. Self-Propag. High-Temp. Synth., 2017, 26, 221–233 CrossRef.
  31. S. H. Park, S. M. Bak, K. H. Kim, J. P. Jegal, S. I. Lee, J. Lee and K. B. Kim, J. Mater. Chem., 2010, 21, 680–686 RSC.
  32. A. A. Zamri, M. Y. Ong, S. Nomanbhay and P. L. Show, Environ. Res., 2021, 197, 111204 CrossRef CAS PubMed.
  33. A. R. Yadav, S. K. Mohite and C. S. Magdum, Asian J. Res. Chem., 2020, 13, 275–278 CrossRef.
  34. A. Kumar, Y. Kuang, Z. Liang and X. Sun, Mater. Today Nano, 2020, 11, 100076 Search PubMed.
  35. W. C. do Nascimento, L. B. Ramo, F. F. da Silva, M. C. U. Araujo, S. I. E. de Andrade and K. M. Bichinho, Spectrochim. Acta, Part A, 2024, 321, 124669 CrossRef CAS PubMed.
  36. P. Chamoli, R. K. Shukla, A. N. Bezbaruah, K. K. Kar and K. K. Raina, Appl. Surf. Sci., 2021, 555, 149663 CrossRef CAS.
  37. J. Chaudhary, G. Tailor, M. Yadav and C. Mehta, Biocatal. Agric. Biotechnol., 2023, 50, 102692 CrossRef CAS.
  38. R. Kaur, A. Mishra and S. Saha, Biocatal. Agric. Biotechnol., 2023, 50, 102723 CrossRef CAS.
  39. K. J. Sreeram, M. Nidhin and B. U. Nair, Bull. Mater. Sci., 2008, 31, 937–942 Search PubMed.
  40. L. Kustov and K. Vikanova, Metals, 2023, 13, 1714 CrossRef CAS.
  41. R. Revathy, J. Joseph, C. Augustine, T. Sajini and B. Mathew, Environ. Sci. Adv., 2022, 1, 491–505 CAS.
  42. J. Xaviour, S. Sreelekshmi, J. Joseph, S. Alfiya and T. Sajini, RSC Sustain., 2024, 2, 3077–3089 RSC.
  43. N. Yadav, R. R. Yadav and K. K. Dey, J. Alloys Compd., 2022, 896, 163073 Search PubMed.
  44. K. P. Theertha, A. Tissamol, S. K. Ashok, R. Revathy and T. Sajini, Chem. Pap., 2023, 78, 1601–1611 CrossRef.
  45. S. A. A. Oliveira, M. Kuznetsova and J. S. Souza, Ceram. Int., 2023, 49, 10795–10806 Search PubMed.
  46. A. H. Jadhav, A. C. Lim, G. M. Thorat, H. S. Jadhav and J. G. Seo, RSC Adv., 2016, 6, 31675–31686 RSC.
  47. S. A. Ogundare, T. O. Adesetan, G. Muungani, V. Moodley, J. F. Amaku, O. Christy Atewolara-Odule, S. T. Yussuf, N. O. Sanyaolu, A. A. Ibikunle, M.-S. Balogun and W. Ewald Van Zyl, Environ. Nanotechnol., Monit. Manage., 2023, 2, 247–256 CAS.
  48. A.-A. Sorescu, A. Nuţă, R.-M. Ion and Ş.-B. Ioana-Raluca, Proc. 4th Int. Virtual Conf. Adv. Sci. Results, 2016, vol. 4, pp. 188–193 Search PubMed.
  49. S. Ahmed, M. Ahmad, B. L. Swami and S. Ikram, J. Adv. Res., 2016, 7, 17–28 Search PubMed.
  50. S. Bhakya, S. Muthukrishnan, M. Sukumaran and M. Muthukumar, Appl. Nanosci., 2016, 6, 755–766 Search PubMed.
  51. S. Wei, Y. Wang, Z. Tang, J. Hu, R. Su, J. Lin, T. Zhou, H. Guo, N. Wang and R. Xu, New J. Chem., 2020, 44, 9304–9312 RSC.
  52. T. Abraham, K. P. Theertha, S. K. Ashok, J. Joseph and T. Sajini, J. Turk. Chem. Soc., Sect. A, 2024, 11, 981–994 CrossRef CAS.
  53. J. Chaudhary, G. Tailor, C. Mehta and M. Yadav, Biocatal. Agric. Biotechnol., 2023, 52, 102812 CrossRef CAS.
  54. D. Manikprabhu and K. Lingappa, Acta Metall. Sin., 2013, 26, 613–617 CrossRef CAS.
  55. S. Joseph and B. Mathew, Spectrochim. Acta, Part A, 2015, 136, 1371–1379 CrossRef CAS PubMed.
  56. L. Castillo-Henríquez, K. Alfaro-Aguilar, J. Ugalde-álvarez, L. Vega-Fernández, G. M. de Oca-Vásquez and J. R. Vega-Baudrit, Nanomaterials, 2020, 10, 1–24 CrossRef PubMed.
  57. S. Zhang, S. Liu, W. Cao, J. Luo, Y. Gu, X. Liu, P. Tan, Z. Wang and J. Pan, J. Colloid Interface Sci., 2024, 675, 1108–1118 CrossRef CAS PubMed.
  58. M. Alaqarbeh, S. F. Adil, T. Ghrear, M. Khan, M. Bouachrine and A. Al-Warthan, Catalysts, 2023, 13, 1343 CrossRef CAS.
  59. K. Seku, S. Sulaiman Hussaini, N. Golla, G. Mangatayaru K, S. M. V. D, S. Rapolu, R. Bandi and B. Reddy G, Mater. Lett., 2020, 278, 128427 CrossRef CAS.
  60. A. K. Mishra, A. K. Nayak, A. K. Das and D. Pradhan, J. Phys. Chem. C, 2018, 122, 11249–11261 CrossRef CAS.
  61. Y. Li, Y. L. Lu, K. Di Wu, D. Z. Zhang, M. Debliquy and C. Zhang, Rare Met., 2021, 40, 1477–1493 CrossRef CAS.
  62. N. Sreeju, A. Rufus and D. Philip, J. Mol. Liq., 2016, 221, 1008–1021 CrossRef CAS.
  63. M. Herlekar, Artic. Int. J. Adv. Res., 2025, 3, 331–345 Search PubMed.
  64. V. Siva, A. Murugan, A. Shameem, S. Thangarasu and S. A. Bahadur, J. Energy Storage, 2022, 48, 103965 CrossRef.
  65. A. Al-Yunus, W. Al-Arjan, H. Traboulsi, R. Schuarca, P. Chando, I. D. Hosein and M. Hessien, Nanomaterials, 2024, 14, 308 Search PubMed.
  66. Z. Zhu, D. Yang and H. Liu, Adv. Powder Technol., 2011, 22, 493–497 CrossRef.
  67. S. Ali, A. S. Sharma, W. Ahmad, M. Zareef, M. M. Hassan, A. Viswadevarayalu, T. Jiao, H. Li and Q. Chen, Crit. Rev. Anal. Chem., 2021, 51, 454–481 Search PubMed.
  68. C. Ragavendran, C. Kamaraj, D. Natarajan, K. A. Al-Ghanim, M. Magesh, M. Nicoletti and M. Govindarajan, Green Process. Synth., 2024, 13, 20230194 CrossRef CAS.
  69. G. Thirumoorthy, B. Balasubramanian, J. A. George, A. Nizam, P. Nagella, N. Srinatha, M. Pappuswamy, A. M. Alanazi, A. Meyyazhagan, K. R. R. Rengasamy and V. Veerappa Lakshmaiah, Sci. Rep., 2024, 14, 1–15 CrossRef PubMed.
  70. R. Araya-Hermosilla, J. Martínez, C. Z. Loyola, S. Ramírez, S. Salazar, C. S. Henry, R. Lavín and N. Silva, Ultrason. Sonochem., 2023, 99, 106545 CrossRef CAS PubMed.
  71. Z. Marková, K. M. H. Šišková, J. Filip, J. Čuda, M. Kolář, K. Šafářová, I. Medřík and R. Zbořil, Environ. Sci. Technol., 2013, 47, 5285–5293 CrossRef PubMed.
  72. A. Lagashetty, S. K. Ganiger, R. K. Preeti and S. Reddy, Biointerface Res. Appl. Chem., 2021, 11, 8087–8095 Search PubMed.
  73. N. G. Dlamini, A. K. Basson and V. S. R. Pullabhotla, Appl. Nano, 2023, 4, 1–24 Search PubMed.
  74. M. A. Malik, S. S. R. Albeladi, S. M. Al-Maaqar, A. A. Alshehri, S. A. Al-Thabaiti, I. Khan and M. R. Kamli, Life, 2023, 13, 653 CrossRef CAS PubMed.
  75. J. Xaviour, S. Sreelekshmi, J. Joseph, S. A. Fathima and T. Sajini, RSC Sustain., 2024, 2, 3077–3089 RSC.
  76. M. Nasrollahzadeh, M. Sajjadi and S. Iravani, Nanomaterials, 2020, 10, 8–14 Search PubMed.
  77. A. Roy, S. Kunwar, U. Bhusal, S. Alghamdi and M. F. N. Azlina, Catalysts, 2023, 13, 321 CrossRef CAS.
  78. A. Kumar Maddheshiya, N. Yadav, S. Yadav, P. Singh Yadav and T. Prasad Yadav, Mater. Lett., 2023, 349, 134735 CrossRef CAS.
  79. Y. Kocak, R. N. E. Tiri, A. Aygun, I. Meydan, N. Bennini, T. Karahan and F. Sen, Bionanoscience, 2024, 14, 93–101 CrossRef.
  80. G. Sharma, S. Bhogal, M. Naushad, Inamuddin, A. Kumar and F. J. Stadler, J. Photochem. Photobiol., A, 2017, 347, 235–243 CrossRef CAS.
  81. M. A. Moradiya, P. K. Khiriya, G. K. Tripathi, P. Bundela and P. S. Khare, Letters in Applied NanoBioScience, 2022, 11, 3102–3110 Search PubMed.
  82. N. Garino, A. Sacco, M. Castellino, J. A. Muñoz-Tabares, M. Armandi, A. Chiodoni and C. F. Pirri, ChemistrySelect, 2016, 1, 3640–3646 Search PubMed.
  83. K. Y. Shih, Y. L. Kuan and E. R. Wang, Materials, 2021, 14, 4577 CrossRef CAS PubMed.
  84. N. Garino, A. Sacco, M. Castellino, J. A. Muñoz-Tabares, A. Chiodoni, V. Agostino, V. Margaria, M. Gerosa, G. Massaglia and M. Quaglio, ACS Appl. Mater. Interfaces, 2016, 8, 4633–4643 CrossRef CAS PubMed.
  85. A. H. Mady, M. L. Baynosa, D. Tuma and J. J. Shim, Appl. Catal., B, 2017, 203, 416–427 CrossRef CAS.
  86. S. Gayathri, M. Kottaisamy and V. Ramakrishnan, AIP Adv., 2015, 5, 127219 CrossRef.
  87. G. Nabiyouni, D. Ghanbari, J. Ghasemi and A. Yousofnejad, J. Nanostruct., 2015, 5, 289–295 Search PubMed.
  88. G. Priyadarshi, N. P. Raval and M. H. Trivedi, Int. J. Biol. Macromol., 2022, 219, 53–67 CrossRef CAS PubMed.
  89. Q. Su, W. Yuan, L. Yao, Y. Wu, J. Zhang and G. Du, Mater. Res. Bull., 2015, 72, 43–49 Search PubMed.
  90. H. W. Kim, H. G. Na, Y. J. Kwon, S. Y. Kang, M. S. Choi, J. H. Bang, P. Wu and S. S. Kim, ACS Appl. Mater. Interfaces, 2017, 9, 31667–31682 CrossRef CAS PubMed.
  91. S. M. Li, L. H. Fu, M. G. Ma, J. F. Zhu, R. C. Sun and F. Xu, Biomass Bioenergy, 2012, 47, 516–521 CrossRef CAS.
  92. T. Lu, L. Pan, H. Li, G. Zhu, T. Lv, X. Liu, Z. Sun, T. Chen and D. H. C. Chua, J. Alloys Compd., 2011, 509, 5488–5492 CrossRef CAS.
  93. W. Wazeer, M. M. Nabil, M. Feteha, M. B. Soliman and A. E. H. B. Kashyout, Sci. Rep., 2022, 12, 1–12 CrossRef PubMed.
  94. S. Krithika and J. Balavijayalakshmi, Mater. Today Proc., 2022, 50, 17–25 CrossRef CAS.
  95. G. Sivakumari, M. Rajarajan and S. Senthilvelan, Res. Chem. Intermed., 2023, 49, 3539–3561 CrossRef CAS.
  96. A. Seza, F. Soleimani, N. Naseri, M. Soltaninejad, S. M. Montazeri, S. K. Sadrnezhaad, M. R. Mohammadi, H. A. Moghadam, M. Forouzandeh and M. H. Amin, Appl. Surf. Sci., 2018, 440, 153–161 CrossRef CAS.
  97. P. Potirak, W. Pecharapa and W. Techitdheera, J. Exp. Nanosci., 2014, 9, 96–105 CrossRef CAS.
  98. P. E. Lokhande, C. Jagtap, V. Kadam, R. Udayabhaskar and S. F. Shaikh, J. Mater. Sci. Mater. Electron., 2024, 35, 1–13 CrossRef.
  99. K. Karthik, S. Dhanuskodi, C. Gobinath and S. Sivaramakrishnan, Spectrochim. Acta, Part A, 2015, 139, 7–12 CrossRef CAS PubMed.
  100. K. Vijayalakshmi and K. Karthick, J. Mater. Sci. Mater. Electron., 2014, 25, 832–836 CrossRef CAS.
  101. J. Y. Lu, Z. Q. Bu, D. Wang, B. He and W. T. Huang, ACS Appl. Eng. Mater., 2024, 2, 1612–1623 CrossRef.
  102. X. Xu, J. Shen, N. Li and M. Ye, J. Alloys Compd., 2014, 616, 58–65 CrossRef CAS.
  103. P. Kumar, S. Dua, R. Kaur, M. Kumar and G. Bhatt, RSC Adv., 2022, 12, 4714–4759 RSC.
  104. S. Ashok Kumar, M. Dheeraj Kumar, M. Saikia, N. Renuga Devi and A. Subramania, Mater. Adv., 2023, 4, 3951–3966 RSC.
  105. G. Rani, Siddharth, R. Ahlawat and H. Kumar, Comments Inorg. Chem., 2024, 45, 1–38 CrossRef.
  106. V. Manikandan and N. Y. Lee, Environ. Res., 2022, 212, 113283 CrossRef CAS PubMed.
  107. Z. Yu, C. Deng, W. Ma, Y. Liu, C. Liu, T. Zhang and H. Xiao, Nanomaterials, 2024, 14, 1827 CrossRef CAS PubMed.
  108. P. M. Olmos-Moya, S. Velazquez-Martinez, C. Pineda-Arellano, J. R. Rangel-Mendez and L. F. Chazaro-Ruiz, Carbon, 2022, 187, 216–229 CrossRef CAS.
  109. P. Yang, Z. Zhu, M. Chen, W. Chen and X. Zhou, Opt. Mater., 2018, 85, 329–336 CrossRef CAS.
  110. P. Kumar, G. Bhatt, R. Kaur, S. Dua and A. Kapoor, Fullerenes, Nanotubes Carbon Nanostruct., 2020, 28, 724–731 CrossRef CAS.
  111. G. Jeong, J. M. Lee, J. ah Lee, J. Praneerad, C. A. Choi, P. Supchocksoonthorn, A. K. Roy, W. S. Chae, P. Paoprasert, M. K. Yeo, G. Murali, S. Y. Park, D. K. Lee and I. In, Appl. Surf. Sci., 2021, 542, 148471 CrossRef CAS.
  112. A. Başoğlu, Ü. Ocak and A. Gümrükçüoğlu, J. Fluoresc., 2020, 30, 515–526 CrossRef PubMed.
  113. H. A. S. Tohamy, M. El-Sakhawy and S. Kamel, Sci. Rep., 2023, 13, 1–11 CrossRef PubMed.
  114. A. Chae, Y. Choi, S. Jo, Nur'aeni, P. Paoprasert, S. Y. Park and I. In, RSC Adv., 2017, 7, 12663–12669 RSC.
  115. M. Li, M. Wang, L. Zhu, Y. Li, Z. Yan, Z. Shen and X. Cao, Appl. Catal., B, 2018, 231, 269–276 CrossRef CAS.
  116. B. K. John, N. John and B. Mathew, Springer Proc. Mater., 2024, vol. 27, pp. 177–183 Search PubMed.
  117. N. Architha, M. Ragupathi, C. Shobana, T. Selvankumar, P. Kumar, Y. S. Lee and R. Kalai Selvan, Environ. Res., 2021, 199, 111263 CrossRef CAS PubMed.
  118. N. R. Pires, C. M. W. Santos, R. R. Sousa, R. C. M. De Paula, P. L. R. Cunha and J. P. A. Feitosa, J. Braz. Chem. Soc., 2015, 26, 1274–1282 CAS.
  119. S. Y. Park, N. Thongsai, A. Chae, S. Jo, E. B. Kang, P. Paoprasert, S. Y. Park and I. In, J. Ind. Eng. Chem., 2017, 47, 329–335 CrossRef.
  120. J. Zhu, C. Wu, Y. Cui, D. Li, Y. Zhang, J. Xu, C. Li, S. Iqbal and M. Cao, Colloids Surf., A, 2021, 623, 126673 CrossRef CAS.
  121. H. Eskalen, S. Uruş, S. Cömertpay, A. H. Kurt and Ş. Özgan, Ind. Crops Prod., 2020, 147, 112209 CrossRef CAS.
  122. J. Prekodravac, B. Vasiljević, Z. Marković, D. Jovanović, D. Kleut, Z. Špitalský, M. Mičušik, M. Danko, D. Bajuk-Bogdanović and B. Todorović-Marković, Ceram. Int., 2019, 45, 17006–17013 CrossRef CAS.
  123. H. Zhang, H. Wang, Y. Wang and B. Xin, Appl. Surf. Sci., 2020, 512, 145751 CrossRef CAS.
  124. J. Hou, H. Li, L. Wang, P. Zhang, T. Zhou, H. Ding and L. Ding, Talanta, 2016, 146, 34–40 CrossRef PubMed.
  125. D. Rodríguez-Padrón, M. Algarra, L. A. C. Tarelho, J. Frade, A. Franco, G. De Miguel, J. Jiménez, E. Rodríguez-Castellón and R. Luque, ACS Sustain. Chem. Eng., 2018, 6, 7200–7205 CrossRef.
  126. S. Ganesan, R. Kalimuthu, T. Kanagaraj, R. Kulandaivelu, R. Nagappan, L. A. Pragasan and V. K. Ponnusamy, Environ. Res., 2022, 206, 112589 Search PubMed.
  127. A. Dager, A. Baliyan, S. Kurosu, T. Maekawa and M. Tachibana, Sci. Rep., 2020, 10, 1–15 CrossRef PubMed.
  128. N. A. Qandeel, A. A. El-Masry, M. Eid, M. A. Moustafa and R. El-Shaheny, Anal. Chim. Acta, 2023, 1237, 340592 CrossRef CAS PubMed.
  129. A. Ahlawat, P. S. Rana and P. R. Solanki, Mater. Lett., 2021, 305, 130830 CrossRef CAS.
  130. Q. Liang, Y. Wang, F. Lin, M. Jiang, P. Li and B. Huang, Anal. Methods, 2017, 9, 3675–3681 RSC.
  131. M. A. A. Hamid, S. H. Elagamy, A. Gamal and F. R. Mansour, Food Anal. Methods, 2024, 17, 739–750 CrossRef.
  132. G. Gyulai, F. Ouanzi, I. Bertóti, M. Mohai, T. Kolonits, K. Horváti and S. Bősze, J. Colloid Interface Sci., 2019, 549, 150–161 CrossRef CAS PubMed.
  133. A. Gholipour, M. Jahanshahi and H. Emadi, J. Clust. Sci., 2024, 35, 237–251 CrossRef CAS.
  134. K. M. Omer, K. H. Hama Aziz, Y. M. Salih, D. I. Tofiq and A. Q. Hassan, New J. Chem., 2019, 43, 689–695 RSC.
  135. S. F. El-Malla, E. A. Elshenawy, S. F. Hammad and F. R. Mansour, Anal. Chim. Acta, 2022, 1197, 339491 CrossRef PubMed.
  136. N. Sharma, I. Sharma and M. K. Bera, J. Fluoresc., 2022, 32, 1039–1049 CrossRef CAS PubMed.
  137. Y. Zhao, S. Zuo and M. Miao, RSC Adv., 2017, 7, 16637–16643 RSC.
  138. H. A. S. Tohamy, M. El-Sakhawy and S. Kamel, RSC Adv., 2024, 14, 25785–25792 RSC.
  139. Z. Ramezani, M. Qorbanpour and N. Rahbar, Colloids Surf., A, 2018, 549, 58–66 CrossRef CAS.
  140. F. M. Abdoon and H. M. Atawy, Mater. Today Proc., 2021, 42, 2141–2149 CrossRef.
  141. M. E. Mahmoud, N. A. Fekry and A. M. Abdelfattah, J. Hazard. Mater., 2020, 397, 122770 CrossRef CAS PubMed.
  142. B. I. Salman, A. I. Hassan, Y. F. Hassan, R. E. Saraya and H. A. Batakoushy, J. Fluoresc., 2023, 33, 1101–1110 CrossRef CAS PubMed.
  143. A. Abdel-Hakim, F. Belal, M. A. Hammad and M. El-Maghrabey, Methods Appl. Fluoresc., 2023, 11, 045007 CrossRef CAS PubMed.
  144. G. Magdy, F. Belal and H. Elmansi, RSC Adv., 2023, 13, 4156–4167 RSC.
  145. C. I. Portela, N. C. S. Vieira, T. R. Brazil, A. S. Giroto, J. B. Gabriel Filho and M. Gonçalves, Environ. Res., 2025, 264, 120311 CrossRef CAS PubMed.
  146. L. Liu, X. Hou, L. Ga, Y. Du, J. Ai and G. Aodeng, Materials Today Quantum, 2025, 5, 100021 CrossRef.
  147. A. Gholipour and S. Rahmani, Fullerenes, Nanotubes Carbon Nanostruct., 2023, 31, 743–751 CrossRef CAS.
  148. Z. Yu, C. Deng, W. Ma, Y. Liu, C. Liu, T. Zhang and H. Xiao, Nanomaterials, 2024, 14, 1827 CrossRef CAS PubMed.
  149. T. u. G. Mir, S. Shukla, A. Q. Malik, J. Singh and D. Kumar, Chem. Pap., 2023, 77, 3641–3649 CrossRef CAS.
  150. Y. Che, H. Pang, H. Li, L. Yang, X. Fu, S. Liu, L. Ding and J. Hou, Talanta, 2019, 196, 442–448 CrossRef CAS PubMed.
  151. R. Kumar and V. K. Bhardwaj, J. Fluoresc., 2024, 1–12 Search PubMed.
  152. M. Ahmed Abdel Hamid, S. H. Elagamy, A. Gamal and F. R. Mansour, Spectrochim. Acta, Part A, 2023, 293, 122440 CrossRef CAS PubMed.
  153. D. Bharathi, B. Siddlingeshwar, R. H. Krishna, V. Singh, N. Kottam, D. D. Divakar and A. A. Alkheraif, J. Fluoresc., 2018, 28, 573–579 CrossRef CAS PubMed.
  154. B. S. Kim, G. H. Oh, Y. Song, Y. S. Lee and S. Kim, Appl. Surf. Sci., 2023, 640, 158301 CrossRef CAS.
  155. A. S. Kamarol Zaman, T. L. Tan, Y. A/P Chowmasundaram, N. Jamaludin, A. R. Sadrolhosseini, U. Rashid and S. A. Rashid, Opt. Mater., 2021, 112, 110801 CrossRef CAS.
  156. A. H. Kamal, R. E. Kannouma, M. A. Hammad and F. R. Mansour, Microchem. J., 2024, 206, 111488 CrossRef CAS.
  157. X. Jiang, Z. Luo, B. Zhang, P. Li, J. Xiao and W. Su, Inorg. Chem. Commun., 2022, 142, 109543 CrossRef CAS.
  158. S. Ingavale, I. Patil, K. Prabakaran and A. Swami, Int. J. Energy Res., 2021, 45, 7366–7379 CrossRef CAS.
  159. H. Wu, P. Cao, W. Li, N. Ni, L. Zhu and X. Zhang, J. Alloys Compd., 2011, 509, 1261–1265 CrossRef CAS.
  160. Z. Li, K. Peng, N. Ji, W. Zhang, W. Tian and Z. Gao, Nanoscale Adv, 2024, 7, 419–432 RSC.
  161. S. K. Kandasamy and K. Kandasamy, Int. J. Electrochem. Sci., 2019, 14, 4718–4729 CrossRef CAS.
  162. Z. Q. Tian, X. L. Wang, H. M. Zhang, B. L. Yi and S. P. Jiang, Electrochem. Commun., 2006, 8, 1158–1162 CrossRef CAS.
  163. W. Lei, L. Wu, W. Huang, Q. Hao, Y. Zhang and X. Xia, J. Mater. Chem. B, 2014, 2, 4324–4330 RSC.
  164. L. Liao, C. Zhang and S. Gong, Macromol. Chem. Phys., 2007, 208, 1301–1309 CrossRef CAS.
  165. A. G. Bekru, L. T. Tufa, O. A. Zelekew, J. Gwak, J. Lee and F. K. Sabir, Crystals, 2023, 13, 133 CrossRef CAS.
  166. N. Liu, X. Wang, W. Xu, H. Hu, J. Liang and J. Qiu, Fuel, 2014, 119, 163–169 CrossRef CAS.
  167. H. Hu, X. Wang, C. Xu, J. Wang, L. Wan, M. Zhang and X. Shang, Fullerenes, Nanotubes Carbon Nanostruct., 2012, 20, 31–40 CrossRef CAS.
  168. A. Saranya, A. Thamer, K. Ramar, A. Priyadharsan, V. Raj, K. Murugan, A. Murad and P. Maheshwaran, J. Photochem. Photobiol., B, 2020, 207, 111885 CrossRef CAS PubMed.
  169. N. Jia, S. M. Li, M. G. Ma, R. C. Sun and L. Zhu, Carbohydr. Res., 2011, 346, 2970–2974 CrossRef CAS PubMed.
  170. A. Uctepe, E. Demir, B. Tekin, B. Dursun, O. Ozturk, O. Sel and R. Demir-Cakan, Solid State Ionics, 2020, 354, 115409 CrossRef CAS.
  171. M. Likhitha, R. R. N. Sailaja, V. S. Priyambika and M. V. Ravibabu, Int. J. Biol. Macromol., 2014, 65, 500–508 CrossRef CAS PubMed.
  172. F. Zhang, Z. Wang, Y. Zhang, Z. Zheng, C. Wang, Y. Du and W. Ye, Int. J. Electrochem. Sci., 2012, 7, 1968–1977 CrossRef CAS.
  173. R. Kumar, K. Kumar, S. Sharma, N. Thakur and N. Thakur, J. Mater. Sci. Mater. Electron., 2023, 34, 1–22 CrossRef.
  174. H. Wang, H. Gao, M. Chen, X. Xu, X. Wang, C. Pan and J. Gao, Appl. Surf. Sci., 2016, 360, 840–848 CrossRef CAS.
  175. T. Preethi, M. P. Pachamuthu, K. Senthil, S. Arulmani, S. Pugalmani and S. Ashokan, J. Mol. Struct., 2024, 1304, 137667 CrossRef CAS.
  176. I. Raya, G. Widjaja, K. Hachem, M. N. Rodin, A. A. Ali, M. M. Kadhim, Y. F. Mustafa, Z. H. Mahmood and S. Aravindhan, J. Nanostruct., 2021, 11, 728–735 CAS.
  177. S. M. Li, N. Jia, M. G. Ma, Z. Zhang, Q. H. Liu and R. C. Sun, Carbohydr. Polym., 2011, 86, 441–447 CrossRef CAS.
  178. Y. R. Lin, F. H. Ko and Y. C. Chang, ACS Appl. Nano Mater., 2024, 7, 16831–16841 CrossRef CAS.
  179. D. Carriazo, J. Patiño, M. C. Gutiérrez, M. L. Ferrer and F. Del Monte, RSC Adv., 2013, 3, 13690–13695 RSC.
  180. K. Dharmalingam, G. Padmavathi, A. B. Kunnumakkara and R. Anandalakshmi, Mater. Sci. Eng. C, 2019, 100, 535–543 CrossRef CAS PubMed.
  181. S. Sahoo, A. Milton, A. Sood, R. Kumar, S. Choi, C. K. Maity and S. S. Han, J. Energy Storage, 2024, 99, 113321 CrossRef.
  182. Y. Wang and J. Y. Lee, J. Power Sources, 2005, 144, 220–225 CrossRef CAS.
  183. B. Govindaraj and M. Sarojadevi, Polym. Adv. Technol., 2018, 29, 1718–1726 CrossRef CAS.
  184. Q. Bu, J. Cai, S. V. Vasudevan, J. Ni and H. Mao, Electrochim. Acta, 2021, 398, 139319 CrossRef CAS.
  185. S. W. Chook, C. H. Chia, S. Zakaria, M. K. Ayob, K. L. Chee, N. M. Huang, H. M. Neoh, H. N. Lim, R. Jamal and R. Rahman, Nanoscale Res. Lett., 2012, 7, 1–7 CrossRef PubMed.
  186. S. Chakraborty, R. Simon, A. Vadakkekara and N. L. Mary, Electrochim. Acta, 2022, 403, 139678 CrossRef CAS.
  187. A. Dubey and C. L. Dube, Nano-Struct. Nano-Objects, 2023, 35, 101017 CrossRef.
  188. Y. Zhang, K. Li, J. Liao, X. Wei and L. Zhang, Appl. Surf. Sci., 2020, 499, 143875 CrossRef CAS.
  189. S. Song, W. Gao, X. Wang, X. Li, D. Liu, Y. Xing and H. Zhang, Dalt. Trans., 2012, 41, 10472–10476 RSC.
  190. Q. Hu, G. Chen, Y. Wang, J. Jin, M. Hao, J. Li, X. Huang and J. Jiang, Nanoscale, 2020, 12, 12336–12345 RSC.
  191. A. Karrab, R. Bensimon, D. Muller, S. Bastide, C. Cachet-Vivier and S. Ammar, Carbon Lett., 2022, 32, 999–1015 CrossRef CAS.
  192. H. E. M. Abdelmoneim, M. A. Wassel, A. S. Elfeky, S. H. Bendary, M. A. Awad, S. S. Salem and S. A. Mahmoud, J. Clust. Sci., 2022, 33, 1119–1128 CrossRef CAS.
  193. I. M. Factori, J. M. Amaral, P. H. Camani, D. S. Rosa, B. A. Lima, M. Brocchi, E. R. da Silva and J. S. Souza, ACS Appl. Nano Mater., 2021, 4, 7371–7383 CrossRef CAS.
  194. S. Jalali, M. Baniadam and M. Maghrebi, Arab. J. Sci. Eng., 2024, 1–12 Search PubMed.
  195. Z. Wang, W. Jia, M. Jiang, C. Chen and Y. Li, Sci. China Mater., 2015, 58, 944–952 CrossRef CAS.
  196. M. A. Karimi, M. Iliyat, M. Atashkadi, M. Ranjbar, A. Habibi-Yangjeh and J. Chinese, Chem. Soc., 2020, 67, 2032–2041 CAS.
  197. R. E. Nimshi, J. J. Vijaya, L. J. Kennedy, P. S. Selvamani, M. Bououdina and P. J. Sophia, Ceram. Int., 2023, 49, 13762–13773 CrossRef CAS.
  198. M. R. Karim, M. M. Alam, M. O. Aijaz, A. M. Asiri, M. A. Dar and M. M. Rahman, Talanta, 2019, 193, 64–69 CrossRef CAS PubMed.
  199. L. Li, X. Zhang, W. Zhang, L. Wang, X. Chen and Y. Gao, Colloids Surf., A, 2014, 457, 134–141 CrossRef CAS.
  200. C. H. Ashok and K. Venkateswara Rao, J. Mater. Sci. Mater. Electron., 2016, 27, 8816–8825 CrossRef CAS.
  201. Y. Yang, E. Liu, J. Fan, X. Hu, W. Hou, F. Wu and Y. Ma, Russ. J. Phys. Chem. A, 2014, 88, 478–483 CrossRef CAS.
  202. F. Rehman, A. Ali, M. Zubair, U. Waheed, R. Khan, A. Yaqoob, I. Shahzadi and M. Siddique, Int. J. Environ. Sci. Technol., 2024, 1–16 Search PubMed.
  203. X. Hangxun, B. W. Zeiger and K. S. Suslick, Chem. Soc. Rev., 2013, 42, 2555–2567 RSC.
  204. O. Kuntyi, L. Bazylyak, A. Kytsya, G. Zozulya and M. Shepida, Biointerface Res. Appl. Chem., 2024, 14, 83 CAS.
  205. R. Seth and A. Meena, Clean Technol. Environ. Policy, 2024, 1–24,  DOI:10.1007/s10098-024-02854-7.
  206. S. Anboo, S. Y. Lau, J. Kansedo, P. S. Yap, T. Hadibarata, J. Jeevanandam and A. H. Kamaruddin, Biotechnol. Bioeng., 2022, 119, 2609–2638 CrossRef CAS PubMed.
  207. S. Tiwari and S. Talreja, J. Pharm. Res. Int., 2022, 34, 74–79 CrossRef.
  208. G. M. Nair, T. Sajini and B. Mathew, Talanta Open, 2022, 5, 100080 CrossRef.
  209. R. A. Sheldon, ACS Sustain. Chem. Eng., 2018, 6, 32–48 CrossRef CAS.
  210. J. Martínez, J. F. Cortés and R. Miranda, Processes, 2022, 10, 1274 CrossRef.
  211. S. B. Patel and D. V. Vasava, ChemistrySelect, 2018, 3, 471–480 CrossRef CAS.
  212. S. Belwal, Mod. Chem., 2013, 1, 22 CrossRef.
  213. V. J. Law and V. J. Law, Am. J. Anal. Chem., 2024, 15, 201–218 CrossRef.
  214. S. Maramai and M. Taddei, Sustain. Org. Synth., 2021, 488–521 Search PubMed.
  215. D. Dallinger and C. O. Kappe, Chem. Rev., 2007, 107, 2563–2591 CrossRef PubMed.
  216. V. Polshettiwar and R. S. Varma, Aqueous Microwave Assisted Chemistry : Synthesis and Catalysis, Royal Society of Chemistry, 2010, vol. 7 Search PubMed.
  217. M. B. Gawande, S. N. Shelke, R. Zboril and R. S. Varma, Acc. Chem. Res., 2014, 47, 1338–1348 Search PubMed.
  218. Y. Wang, Q. Hou, M. Ju and W. Li, Nanomaterials, 2019, 9, 647 CrossRef PubMed.
  219. M. G. Ma, J. F. Zhu, Y. J. Zhu and R. C. Sun, Chem. An Asian J., 2014, 9, 2378–2391 CrossRef PubMed.
  220. S. Rabieh, M. Bagheri, M. Heydari and E. Badiei, Mater. Sci. Semicond. Process., 2014, 26, 244–250 CrossRef CAS.
  221. R. Shahid, M. Gorlov, R. El-Sayed, M. S. Toprak, A. Sugunan, L. Kloo and M. Muhammed, Mater. Lett., 2012, 89, 316–319 CrossRef.
  222. T. N. Pham-Truong, T. Petenzi, C. Ranjan, H. Randriamahazaka and J. Ghilane, Carbon, 2018, 130, 544–552 CrossRef CAS.
  223. A. Pal, S. Shah and S. Devi, Mater. Chem. Phys., 2009, 114, 530–532 CrossRef CAS.
  224. R. Sunil, J. Joseph and T. Sajini, South African J. Chem., 2024, 78, 192–203 CAS.
  225. D. A. Patiño-Ruiz, S. I. Meramo-Hurtado, Á. D. González-Delgado and A. Herrera, ACS Omega, 2021, 6, 12410–12423 CrossRef PubMed.
  226. B. S. Ramadan, Y. G. Wibowo, D. Anwar and A. T. Maryani, Glob. Nest J., 2024, 26(7) DOI:10.30955/gnj.06216.
  227. A. Bafana, S. V. Kumar, S. Temizel-Sekeryan, S. A. Dahoumane, L. Haselbach and C. S. Jeffryes, Sci. Total Environ., 2018, 636, 936–943 CrossRef CAS PubMed.
  228. D. L. de Melo, A. Kendall, J. T. DeJong, A. Säynäjoki, J. Heinonen, S. Junnila, E. Zahra, M. Manish, J. Baltrusaitis, J. Munuera, L. Britnell, C. Santoro, R. Cuéllar-Franca and C. Casiraghi, 2D Mater., 2021, 9, 012002 Search PubMed.
  229. M. F. Ordoñez, S. Biella, E. Falletta and C. L. Bianchi, Chem. Eng. J. Adv., 2023, 16, 100575 CrossRef.
  230. V. Ntouros, I. Kousis, D. Papadaki, A. L. Pisello and M. N. Assimakopoulos, Energies, 2021, 14, 4998 CrossRef CAS.
  231. D. Papadaki, S. Foteinis, G. H. Mhlongo, S. S. Nkosi, D. E. Motaung, S. S. Ray, T. Tsoutsos and G. Kiriakidis, Sci. Total Environ., 2017, 586, 566–575 CrossRef CAS PubMed.
  232. R. Sendão, M. del, V. M. de Yuso, M. Algarra, J. C. G. Esteves da Silva and L. Pinto da Silva, J. Clean. Prod., 2020, 254, 120080 CrossRef.
  233. S. Fernandes, J. C. G. E. da Silva and L. P. da Silva, Materials, 2022, 15, 3446 CrossRef PubMed.
  234. P. Ubolsook, K. Khamfong, P. Jansanthea, W. Chomkitichai, J. Ketwaraporn, S. Tangjuank, C. Wansao, A. Wanaek, A. Wannawek, Y. Keereeta, S. Kuimalee and P. Pookmanee, Emergent Mater., 2025, 1–21 Search PubMed.
  235. P. N. Omo-Okoro, C. E. Maepa, A. P. Daso and J. O. Okonkwo, Waste Biomass Valorization, 2020, 11, 2247–2259 Search PubMed.
  236. W. L. Ang, Q. A. Alqasem and A. W. Mohammad, IOP Conf. Ser.:Mater. Sci. Eng., 2021, 1195, 012008 Search PubMed.
  237. R. Revathy, J. Joseph, C. Augustine and T. Sajini, Environ. Sci. Adv., 2022, 1, 491–505 CAS.
  238. M. Susan Punnoose, D. Bijimol and B. Mathew, Environ. Nanotechnol., Monit. Manage., 2021, 16, 100525 CAS.
  239. S. Vijayaram, H. Razafindralambo, Y. Z. Sun, S. Vasantharaj, H. Ghafarifarsani, S. H. Hoseinifar and M. Raeeszadeh, Biol. Trace Elem. Res., 2023, 202, 360–386 CrossRef PubMed.
  240. T. Chopra and R. Parkesh, ACS Omega, 2024, 9, 4555–4571 CrossRef CAS PubMed.
  241. R. R. Pathan, M. S. Shaikh, I. J. Syed, M. A. Qureshi, P. N. Acharya, M. K. Shirsat, Aquil-ur-RahimSiddiqui, M. H. Dehghan and R. S. Moon, Nano-Struct. Nano-Objects, 2024, 37, 101116 CrossRef CAS.
  242. M. J. Williams, E. Sánchez, E. R. Aluri, F. J. Douglas, D. A. Maclaren, O. M. Collins, E. J. Cussen, J. D. Budge, L. C. Sanders, M. Michaelis, C. M. Smales, J. Cinatl, S. Lorrio, D. Krueger, R. T. M. De Rosales and S. A. Corr, RSC Adv., 2016, 6, 83520–83528 RSC.
  243. D. S. Rana, S. Kalia, R. Kumar, N. Thakur, D. Singh and R. K. Singh, Mater. Chem. Phys., 2022, 287, 126283 CrossRef.
  244. M. Elkady, H. Shokry and H. Hamad, Chem. Eng. Technol., 2018, 41, 553–562 CrossRef.
  245. D. Ozyurt, M. Al Kobaisi, R. K. Hocking and B. Fox, Carbon Trends, 2023, 12, 100276 CrossRef.
  246. L. Cui, K. N. Hui, K. S. Hui, S. K. Lee, W. Zhou, Z. P. Wan and C. N. H. Thuc, Mater. Lett., 2012, 75, 175–178 CrossRef CAS.
  247. L. Li, Z. Guo, A. Du and H. Liu, J. Mater. Chem., 2012, 22, 3600–3605 RSC.
  248. J. Y. Lu, Z. Q. Bu, Y. Q. Lei, D. Wang, B. He, J. Wang and W. T. Huang, J. Mol. Liq., 2024, 409, 125503 CrossRef CAS.
  249. S. Das, S. Senapati, R. Ganesan and R. Naik, ACS Appl. Nano Mater., 2024, 7, 1–9 Search PubMed.
  250. P. Porrawatkul, R. Pimsen, A. Kuyyogsuy, N. Teppaya, A. Noypha, S. Chanthai and P. Nuengmatcha, RSC Adv., 2022, 12, 15008–15019 RSC.
  251. M. B. Patil, S. B. Rajamani, S. N. Mathad, A. Y. Patil, M. A. Hussain, H. S. Alorfii, A. Khan, A. M. Asiri, I. Khan and M. Puttegowda, J. Mater. Res. Technol., 2022, 20, 3537–3548 CrossRef CAS.
  252. B. S. Kaith, Sukriti, J. Sharma, T. Kaur, S. Sethi, U. Shanker and V. Jassal, Iran. Polym. J., 2016, 25, 787–797 CrossRef CAS.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.