Next Article in Journal
Biomass Pyrolysis-Derived Biochar: A Versatile Precursor for Graphene Synthesis
Previous Article in Journal
Moisture Transport in Loose Fibrous Insulations under Steady-State Boundary Conditions
Previous Article in Special Issue
Review of the Developments and Difficulties in Inorganic Solid-State Electrolytes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Brief Review of MoO3 and MoO3-Based Materials and Recent Technological Applications in Gas Sensors, Lithium-Ion Batteries, Adsorption, and Photocatalysis

by
Mário Gomes da Silva Júnior
*,
Luis Carlos Costa Arzuza
,
Herbet Bezerra Sales
,
Rosiane Maria da Costa Farias
,
Gelmires de Araújo Neves
,
Hélio de Lucena Lira
and
Romualdo Rodrigues Menezes
*
Laboratory of Materials Technology (LTM), Department of Materials Engineering, Federal University of Campina Grande (UFCG), Av. Aprígio Veloso 882, Campina Grande 58429-900, PB, Brazil
*
Authors to whom correspondence should be addressed.
Materials 2023, 16(24), 7657; https://doi.org/10.3390/ma16247657
Submission received: 3 November 2023 / Revised: 1 December 2023 / Accepted: 7 December 2023 / Published: 15 December 2023

Abstract

:
Molybdenum trioxide is an abundant natural, low-cost, and environmentally friendly material that has gained considerable attention from many researchers in a variety of high-impact applications. It is an attractive inorganic oxide that has been widely studied because of its layered structure, which results in intercalation ability through tetrahedral/octahedral holes and extension channels and leads to superior charge transfer. Shape-related properties such as high specific capacities, the presence of exposed active sites on the oxygen-rich structure, and its natural tendency to oxygen vacancy that leads to a high ionic conductivity are also attractive to technological applications. Due to its chemistry with multiple valence states, high thermal and chemical stability, high reduction potential, and electrochemical activity, many studies have focused on the development of molybdenum oxide-based systems in the last few years. Thus, this article aims to briefly review the latest advances in technological applications of MoO3 and MoO3-based materials in gas sensors, lithium-ion batteries, and water pollution treatment using adsorption and photocatalysis techniques, presenting the most relevant and new information on heterostructures, metal doping, and non-stoichiometric MoO3−x.

1. Introduction

Molybdenum oxides are versatile compounds that have attracted the attention of many researchers worldwide since they are one of the most adaptable and functional optical and electronic oxides [1]. Among these oxides, MoO3 is an abundant natural and sustainable material that has gained considerable attention due to its chemistry related to multivalent Mo, high thermal and chemical stability, and redox chemistry resulting from its high reduction potential and electrochemical activity [2,3,4,5]. Due to its large band gap and optical and electrical properties, MoO3 is considered a potential and appropriate material for various technological applications in the fields of photocatalysis, adsorption, gas sensing, battery electrodes as anodic and cathodic materials, recording material, electrochromic and photochromic materials, organic solar cells, and organic light-emitting diodes as a buffer layer and as a catalyst for the electrochemical reduction in several ions [6,7,8]. Interestingly, MoO3 and MoO3−x oxides have been used as hole injection/extraction interlayers in organic solar cells in order to modify the anode contact due to their high work function and their ability to exchange charges with several semiconductors, which requires the energy-level alignment of the materials. As a result, there is a reduction in the barrier for hole extraction from the organic photoactive layer to the metallic anode through the formation of a large interfacial dipole [9,10,11]. MoO3-based materials are also widely used in various industries as display devices and smart windows and are an excellent anti-friction and anti-wear agent as a lubricant additive [2,12,13]. Its absorption band is close to the maximum of human eye sensitivity. Furthermore, it also exhibits a better electrochromic response than other oxides when colored, owing to more intense and regular light absorption [14]. This high applicability is confirmed by verifying the many studies carried out in different technological segments like nanotechnology, energy fuels, electrochemistry, environmental science, optics, etc. An increasing trend in the last 20 years in papers directly related to MoO3 and MoO3-based systems has been verified (Figure 1).
Molybdenum has oxidation states ranging from +2 to +6. However, molybdenum oxides are primarily found in two forms: Mo (IV) oxide and Mo (VI) oxide [15]. Being the most usual, MoO3 is characterized by three different crystalline polymorphs: thermodynamically stable orthorhombic (α-MoO3), metastable monoclinic (β-MoO3), and low-temperature metastable hexagonal (h-MoO3) [2,15,16,17,18]. The different MoO3 crystalline structures are illustrated in Figure 2. In particular, α-MoO3 is an n-type indirect semiconductor with a wide band gap (2.8–3.6 eV) depending on its oxygen vacancy content [15,18,19]. It has high ionic conductivity due to its natural tendency toward oxygen vacancy formation and high stability in the air [5,7]. Moreover, one of the distinguishing characteristics of α-MoO3 is its intrinsic layered nature, which allows large quantities of positive ions to be accommodated in its tetrahedral and octahedral holes and extension channels, allowing bandgap manipulations. Furthermore, the layered structure also leads to superior charge transfer [5,20].
The properties and oxidation states of these phases change with temperature and energy supply, making the fabrication of various molybdenum oxide-based materials possible [21]. Besides MoO3, the crystallographic phases of molybdenum oxides include MoO2, Mo8O23, and Mo4O11 [21,22]. To engineer and control the electronic states of molybdenum oxides, oxidation state variations obtained from oxygen vacancy formation and dopant placement are used to enable crystal structure and morphology manipulation [1]. In particular, oxygen vacancy insertion is one of the best ways to control the doping levels in molybdenum oxides. This is because different amounts of oxygen vacancies can be made with high precision, controlling the change from semiconducting to metallic properties. Specifically, the electronic and optical properties of the MoO3 semiconducting oxide are significantly changed by the oxygen vacancies insertion. Regarding the non-stoichiometric MoO3−x, when 0 < x < 0.125, this oxide exhibits semiconducting properties, while 0.125 < x < 1 exhibits quasimetallic properties, and for x = 1, it creates semiconducting MoO2 with a reduced bandgap or semimetallic [1]. To comprehend the reactions that take place on the MoO3−x surfaces, it is essential to understand the stability, shape, and replenishment techniques of vacancy creation [1]. One efficient technique to create a gap state in MoO3 is to anneal it in reducing gas, which removes the oxygen ions from its rigid metal-oxygen network [23]. The creation and control of oxygen vacancies can also be achieved by H+ intercalating into molybdenum oxide, forming the hydrogen molybdenum bronzes HxMox5+Mo1x6+O3 [24]. Hydrogenous intercalation does not considerably modify the crystalline structure, but there are small changes in volume, lattice distortion, and ordering of H (right panel, Figure 3). Furthermore, due to the proximity between hydrogen and oxygen, water is formed, changing the metal/oxygen ratio and, consequently, the electronic structure. The left panel of Figure 3 shows schematically the effect on the band structure of H+ intercaled into MoO3.
The semiconducting MoO2 has a monoclinic lattice system and can be made by removing the stoichiometric oxygen from MoO3 in a C or H2 reduction atmosphere. Because of its increased free electrons associated with oxygen vacancy richness, MoO2 has stronger metallic properties than semiconducting properties, with a significantly reduced bandgap [1]. The reaction that reduces MoO3 to MoO2 can be explained in two ways. In the first one, the MoO3 reduction initiates with the creation of oxygen vacancies arranged as shear planes as a consequence of the crystallographic shear that aggregates into discs. As the concentration of the shear planes increases, they order and form the Magnelli phases (MonO3n−1) as an intermediate product, an orthorhombic Mo4O11 [25,26,27]. At higher temperatures, the structure is then converted to MoO2 due to the formation of a critical oxygen vacancy concentration [25]. Thereby, the rate-controlling steps of MoO3 to Mo4O11 and Mo4O11 to MoO2 are the interfacial chemical reactions that obey the nucleation and growth models and the temperature-dependent processes that obey the diffusion models, respectively [27]. The second explanation is that the MoO3 reduction initiates with the hydrogen adsorption onto terminal oxygen atoms and in the layer at oxygen atoms of corner-sharing MoO6 units along the [1 0 0] direction [25]. The MoO3 reduction to MoO2 by hydrogen is an exothermic reaction with the release of heat and consequent local temperature increase [27].
On the other hand, the morphology generally affects MoO3 performance, and this is because MoO3 has a unique structure. For the most part, the different morphologies of MoO3 and MoO3-based materials determine their applications. For example, MoO3 nanorods usually have porous structures that will improve charge and mass migration, which can be used for energy storage devices. MoO3 nanotubes are regarded as one of the most promising architectures for ion batteries and gas sensors because of their large specific surface area, low density, and superior transport. Nanoflowers and nanospheres can make the surface area bigger and give it more active sites, which can improve its performance in areas like sensing, energy storage conversion, and catalysis [28,29,30].
This article aims to briefly review and highlight the main interesting and applicable properties and the latest advances in technological applications of MoO3 and MoO3-based materials in gas sensors, as anode and cathode materials for lithium-ion batteries, and water pollution treatment using adsorption and photocatalysis techniques, presenting the most relevant and new information on heterostructures, metal doping, and non-stoichiometric MoO3−x.

2. Air Pollution and Gas Sensing Applications

Air pollution is considered one of the major 21st-century challenges due to the rapid growth of industries and automobiles, which cause high air pollutant emissions [31]. Among the various pollutants, volatile organic compounds (VOCs) are mostly obtained from exhaust gases, fuel combustion, and transportation. People can get headaches, nausea, convulsions, and commas if exposed to a certain VOC concentration, and many carcinogens can also harm the liver, kidneys, brain, and neurological system [32]. For this reason, there is a growing need to design and fabricate sensing devices with high response and selectivity, low fabrication costs, and a flexible mechanical nature to control air quality [31]. Smart gas sensors are currently widely used in air quality monitoring (humidity or oxygen sensing in car cabins or houses), biosensors for diagnosing a variety of diseases, and leak detection of harmful and explosive gas emissions in industries, which may lead to severe accidents and health hazards [33].
Generally, a gas sensor is a device or chemical sensor that typically consists of a transducer with an active layer that transforms the analyte species information into electrical signals from many detection principles, such as changes in conductivity, resistance, absorbance, etc., in the presence of gas molecules. As electrical properties change, the gas analyte interacts as a donor or an acceptor of charge carriers, changing the material’s resistivity. This resistance change is related to the type of majority carriers present and also to the oxidizing or reducing nature of the gas [33]. The receptor function of the gas sensing material, its converting function, and its utility function influence gas-sensing properties, particularly sensitivity. The receptor function is concerned with the oxide capacity of the surface to interact with the detecting gas molecules, while the transforming function is concerned with the signal conversion to an electric signal, which is Induced by the chemical interaction between the oxide surface and the detecting gas. The utility function is associated with the availability of the inner oxide grains for the molecules of the detecting gas [34].
Among the existing gas sensors, the metal oxide-based ones, known as resistive (Chemi-resistors) or hot-wire sensors, present interesting advantages in their good reliability and simple implementation for real-time monitoring of various gases as VOCs [35,36]. These advantages are attributed to their unique merits, such as high structural stability, portability, adjustable electrical properties, reduced grain size, surface morphologies, easy production, simple operation, and ease of system integration [33,37,38,39]. For this reason, a lot of research efforts have been invested in this kind of gas sensor since the first fundamental idea demonstration in 1962 [39]. This sensor type shows itself to be sensitive to environmental chemical reactions, and due to its stability to operate in harmful environments, it can compete with other sensor types in terms of reliability, sensitivity, and accuracy, having received much more attention since the second half of the 20th century [36].
The molybdenum trioxide (MoO3), as a wide band gap n-type and layered structured semiconductor, has been extensively investigated and applied in gas sensors, where it has been found to be sensitive to different gases such as H2S [40], NO2 [31,41,42], NO [31], H2 [36], CO [43], CO2 [44], C2H5OH [34,35,45,46], NH3 [47], trimethylamine [48,49], acetone, 1-butylamine [37], triethylamine [38,50,51], formaldehyde [52], water/humidity sensor [53] and glucose [54], with detection limits often in ppm range [31,35,36]. A reaction mechanism generality for sensing the gases is shown in Figure 4. Initially, the oxygen molecules are absorbed at the surface of MoO3, adding the depletion region, i.e., the electron-depleted space charge region in which the absorbed oxygen molecule captures electrons from the MoO3-conduction band, as a result of its high electron affinity (Figure 4, left panel). Then, in the presence of a gas sensing, for instance H2, the previously absorbed oxygen molecules react with the H2 gas and subsequently release the captured electrons from the MoO3 conduction band. This reduces the depletion region and increases MoO3 conductivity (Figure 4, right panel).
According to Figure 5, there has been an increasing trend in the last 20 years in papers directly related to MoO3 and MoO3-based systems applied to gas sensors. A summary of recent research on MoO3 and MoO3-based systems in gas sensing applications in the last seven years is presented in Table 1.
Moreover, this oxide has gained significant interest as an electrode material for electrochemical sensors owing to its high electrochemical activity, high dielectric constant, and good gas molecule adsorption [55]. Because of their charge carrier concentration modulation in response to oxidizing and reducing gases, molybdenum oxides are employed in conductometric gas sensing [1]. The performance of these MoO3-based gas sensors depends on gas type sensing (oxidizing or reducing gases), as different reaction mechanisms occur between the molybdenum oxide surface and the target gas molecule, leading to a change in the conductivity of MoO3 (Figure 6). Interestingly, the MoO3 sensing mechanism is considered to be different from those of other common n-type semiconductors (e.g., SnO2 or ZnO). It is proposed to be a lattice oxygen reaction (catalyst property) rather than a traditional surface-chemisorbed oxygen mechanism due to its crystalline-layered structure [37,56]. The lattice oxygen reaction mechanism indicates that the decrease in resistance is related to the free electrons created from the oxygen vacancies by the partial reduction of Mo6+ to Mo5+ ions via the catalytic oxidation reaction between the reducing agents and the lattice oxygen [37]. In this process, shear structures are formed by removing lattice oxygen at the active oxide surface, which is related to the ease of rearrangement of polyhedral [56].
Despite its attractive properties, the unsatisfactory MoO3 electrical conductivity limits its electrochemical performance [55]. To improve this, MoO3 functional and surface modifications, morphological and sizing adjustments by crystal growth interference exposing different crystal planes, doping elements, loading noble metals, and composite synthesis have been considered as interesting options [43,57,58]. The chemisorption activation energy can be effectively decreased with suitable metal doping [32]. These metal elements work as catalytic activity centers that change the way electrons move inside the sensor and make it better at detecting gases, obtaining a considerable response, remarkable selectivity, reliable stability, and a rapid recovery rate [32,38,43]. The heterostructures interfere with properties such as carrier concentration, grain boundary barrier, depletion layer, and energy band, improving gas sensing performance [32]. Several types of research have then been performed using nanocomposites applied to different gases, for instance, carbon monoxide, triethylamine, trimethylamine, formaldehyde, 1-butylamine, hydrogen, benzene, toluene, xylene, glucose, ethyl acetate, and ethanol, like Cr-MoO3 [38], Zn-MoO3 [43], Pt-MoO3 [36,52], Au-MoO3 [37,49,59], TiO2-MoO3 [34], MoO3-ZnO [46], SnS-MoO3 [51], NiO-MoO3 [60], PVP-MoO3 [54], and MoO3 nanowires [61].
Particularly, the satisfactory high reactivity and thermodynamic stability of some reducing agents make the α-phase MoO3 a standout option for this application [57]. The layered structure associated with α-MoO3 increases the Mo5+ pentavalent ions, which have a strong affinity for oxygen [30]. The electrical conductance of non-stoichiometric MoO3−x is higher than that of stoichiometric metal oxide due to enhanced charge carrier mobility associated with the presence of low-valent Mo ions, which is the sensing principle of MoO3 [37,56]. One way to synthesize non-stoichiometric MoO3−x materials and consequently improve the selectivity and sensing performance of these sensors is to introduce oxygen vacancies into the lattice, causing structural defects. The presence of these vacancies plays a significant role in the surface charge distribution under thermal excitation. Then, the band gap can be adjusted by these lattice vacancy insertions, creating gap states. These oxygen vacancies also act as an adsorbing center with a high affinity toward gases at high temperatures. Moreover, the environmental oxygen molecules adsorb spontaneously on the surface of vacancy sites, leading to the formation of more active oxygen species that trap free electrons next to the surface area. At lower temperatures, these oxygen vacancies are not reactive, leading to high resistance in the channel. Otherwise, higher operating temperatures allow abundant thermal energy to exceed the activation energy barrier to accomplish the surface reaction, resulting in an increasing free charge carrier density close to the conduction [56].
The presence of oxygen vacancies can often cause several interface and surface physicochemical changes, including the formation of a new donor level in the forbidden band and an increase in the concentration of available carriers [57]. The electrical conduction of MoO3 is mostly dependent on the free electrons available in its conduction bands. The defect concentration, such as an oxygen deficit, significantly impacts the free-electron intensities in these materials [1]. Additionally, under the influence of oxygen vacancies, the surface lattice structure can change from a stable state to a metastable state, resulting in the formation of an unsaturated coordination metal atom at the vacancy as the center of free electrons and opening up a new pathway for the interaction of electrons with target gas molecules. The amount of surface chemisorbed oxygen increases as a result of plentiful oxygen vacancies acting as active sites that can also provide abundant unpaired electrons acting as electron donors, boosting the adsorption activity [57].
Table 1. MoO3 and MoO3-based systems used in gas sensing applications. Optimal sensing: ratio between the resistance of the target gas and the reference gas (N2). Optimal temperature: working temperature.
Table 1. MoO3 and MoO3-based systems used in gas sensing applications. Optimal sensing: ratio between the resistance of the target gas and the reference gas (N2). Optimal temperature: working temperature.
MaterialMorphologySynthesis MethodWorking Temp. (°C)Gas (Conc.)Optimal
Sens. (%)
Optimal Temp. (°C)Reference
α-MoO3Thin FilmsPulsed laser deposition 100 and 200NO2
(10 ppm)
25100[31]
α-MoO3NanobeltsHydrothermal(250–400)C2H6O (800 ppm)173300[35]
α-MoO3NanobeltsHydrothermal(100–400)1-butylamine (100 ppm)~90340[37]
α-MoO3NanorodsHydrothermal(100–350)Triethylamine (100 ppm)~115200[38]
α-MoO3Nanoflowers/nanosheetsIn situ oxidation(100–350)H2S (10 ppm)~38250[40]
α-MoO3Large-sizedVapor-phase transport (50–125)NO2 (10 ppb)2.3100[42]
α-MoO3Hierarchical
microflower
Thiourea-assisted hydrothermal (120–380)CO (50 ppm)8.02260[43]
α-MoO3NanoparticlesThermal evaporation (150–250)CO2 (150 ppm)15250[44]
α-MoO3Hierarchical nanofiber-assembledHydrothermal(50–350)C2H6O (400 ppm)24300[45]
α-MoO3Hierarchical nanosheet-assembledHydrothermal(50–350)C2H6O (400 ppm)32300[45]
α-MoO3NanorodsThermal evaporation under vacuum RT aNH3 (100 ppm)886RT a[47]
α-MoO3Porous nanosheetsSolvothermal50, 90, 133, 172, and 217Trimethylamine (10 ppm)51.47133[48]
α-MoO3NanorodsHydrothermal(100–350)Triethylamine (100 ppm)153.36200[50]
α-MoO3NanorodsHydrothermal250Triethylamine (50 ppm)73.46250[51]
α-MoO3NanowiresHydrothermal(200–320)Triethylamine (500 ppm)~4150280[61]
h-MoO3NanosheetsDispersionRT aNH3 (0.3 ppm)3.78RT a[39]
α-MoO3−xThin filmsChemical vapor deposition(160–270)NO2 (10 ppm)56250[56]
α-MoO3−xThin filmsChemical vapor deposition(160–270)H2S (10 ppm)18250[56]
Pt-α- MoO3NanoparticlesDispersion(170–300)H2 (500 ppm)18260[36]
Au-α-MoO3Nanoparticles/NanobeltsHydrothermal(100–400)1-butylamine (100 ppm)~300240[37]
Zn-MoO3Hierarchical
microflower
Thiourea-assisted hydrothermal (120–380)CO (50 ppm)31.23240[43]
Au-MoO3Nanoparticles/nanobeltsHydrothermal + Dispersion(180–300)Trimethylamine (50 ppm)70280[49]
Pt-MoO3NanobeltsHydrothermal + DispersionRT aHCHO (200 ppm)39.3RT a[52]
Au-MoO3Nanoparticles/hollow spheresSurfactant-modified approach(217–330)Toluene (100 ppm)17.5250[59]
Au-MoO3Nanoparticles/hollow spheresSurfactant-modified approach(217–330)Xylene (100 ppm)22.1250[59]
Cr2O3-α-MoO3NanorodsHydrothermal(100–350)Triethylamine (100 ppm)150.25200[38]
α-MoO3-ZnONanoparticles/nanobeltsHydrothermal(100–300)C2H6O (400 ppm)19250[46]
α-MoO3-SnSNanotubesHydrothermal250Triethylamine (50 ppm)36.06250[51]
NiO-α-MoO3Core-shell nanorodsHydrothermal + Dispersion(150–350)(100 ppm)34.91250[60]
h-BN-α-MoO3Nanowires/2D materialHigh-temperature pyrolysis(200–320)Triethylamine (500 ppm)8616260[61]
a Room Temperature.
Compared to other metal oxides, α-MoO3 is considered a potential material for detecting triethylamine (TEA), since it is an excellent catalyst for dissociating the carbon-nitrogen bonds of organic amines and specifically interacts with TEA molecules due to strong interactions between acidic α-MoO3 surfaces and basic TEA and the lower activation energy for the carbon-nitrogen bonds compared to other chemical bonds. The Cr-doping into the α-MoO3 lattice promoted the Mo5+ existence due to the similar ionic radii of Mo+6 and Cr+3, resulting in higher responses to TEA [38]. The 2D α-MoO3−x sensors also presented dual gas sensing characteristics, with excellent sensitivity and selectivity toward H2S (10 ppm) and NO2 (10 ppm) at 250 °C, compared to other gases such as H2 (1000 ppm), CO (1200 ppm), CO2 (1000 ppm), and CH4 (10,000 ppm). The highly electrophilic nature of NO2 was shown to contribute to its strong oxidizing properties and to be extremely reactive toward the intrinsic oxygen vacancy sites on the α-MoO3−x surface. Upon H2S exposure, the surface oxygen vacancy sites of MoO3−x showed an interaction with H2S gas molecules and released charge carriers, resulting in defect states within the energy gap. Hence, the sensor presented an n-type behavior of MoO3−x toward a reducing gas [56].
When compared to alcohols (methanol, ethanol, and n-butanol), acetone, and pyridine, MoO3 and Au-MoO3 nanocomposites were very selective for 1-butylamine. The unique layered structure of MoO3 and the type of connection within the 1-butylamine are responsible for its strong selectivity. Among these gases (1-butylamine, methanol, ethanol, n-butanol, and pyridine), both n-butanol and 1-butylamine molecules have a butyl group, which is a relatively strong electron-donating group compared with methyl and ethyl groups. The nitrogen atom in 1-butylamine and the oxygen atom in n-butanol have lone pair electrons that act as electron donors to form bonds. Such structure is preferentially absorbed onto Mo ions that are considered Lewis-acid sites (see Figure 7). The Au-MoO3 nanocomposites also showed a high response to NH3 and NO2, associated with the MoO3 and N atom interactions [37].
Moreover, the α-MoO3 surface showed to be beneficial for the adsorption of H2O molecules because of the pentacoordinated Mo atoms introduced by the interchain interactions breaking, which possessed the Lewis acid character of an electron pair acceptor. The adsorption of H2O molecules on these Mo sites occurs mainly to balance the electrostatic force around the Mo atoms, allowing the obtaining of a robust humidity sensor presenting a fast response and high sensitivity within the wide sensing range of 10–90% relative humidity (RH) [53].

3. Lithium-Ion Batteries

Lithium-ion batteries (LIBs) have become one of the most important and practical energy storage systems for power sources due to their power density, higher energy, longer lifespan, high operating voltage, excellent rechargeability, stable cycle performance, no memory effects, considerable environmental friendliness, and beneficial safety [24,62,63,64,65,66,67]. The LIBs have been researched for use in a variety of electronic devices, including laptops and cell phones, as well as electric and hybrid cars [24,62,63]. Two solid-state electrodes (the cathode and anode), an electrolyte containing Lithium ions, a spacer polymer, and two current collectors typically make up Lithium batteries (Figure 8). Each component of the battery is crucial; for instance, the separator prevents a short circuit inside the cell, and the electrode material determines how much energy they can store. Thus, research on electrode material property election is fundamental to the storage efficiency of the battery.
Nowadays, graphite dominates as an anode material in LIB because of its advantages of great abundance and low cost. However, its lower and limited theoretical Li-storage capacity (372 mA·h·g−1) and slower discharge potential are insufficient to fulfill the increasing energy-consuming demands [24,63]. Alternatively, molybdenum oxides have been researched as anode materials for lithium-ion batteries, benefiting from their higher specific capacities, acceptable cost, non-toxicity, and high stability [24,62,68]. Interestingly, MoO3 is also a potential Li-ion battery material that can function as a cathode [69,70,71,72]. For example, Wang and co-workers used an ammonolysis process to modify the crystal structure and consequently enhance the electrochemical performance of MoO3 nanobelts when tested as cathodes for Li-ion batteries [70]. The modified nanobelts exhibited a capacity of 250 A·h/kg in the potential window of 1.5–3.5 V. The unfavorable operating voltage of reduced Mo oxides makes them possible for use as either cathodes or anodes. Typical redox voltages for Mo oxides lie in a regime that hinders maximizing energy density when they are paired with higher-voltage cathodes or lower-voltage anodes [73]. Due to the wide range of oxidation states (from +6 to +2), they are promising as both positive (cathode) and negative (anode) electrodes in electrochemical cells [74].
One of the most important and interesting properties of molybdenum oxides is their intercalation ability with a wide range of ion sites, which is possible if the host material layers or tunnels have weak binding [75]. This is a consequence of the weak Van Der Waals force between the stacked sheets along the [0 1 0] direction [68]. The α-MoO3 characteristic layer structure forms tetrahedral and octahedral holes and extension channels, leading to more diffusing channels and embedding sites for Li+ ions [75].
Molybdenum oxides can experience redox reactions that are highly reversible, such as Mo6+/Mo4+. Theoretically, these reactions produce a high energy density of 745 W·h·kg−1 because they permit numerous electron exchanges per molybdenum redox center and can load up to 1.5 Li per molybdenum atom [1,68]. Additionally, the potential Li-storage capacities of the MoO3 and MoO2 presented have high theoretical specifications: 1117 mA·h·g−1 and 838 mA·h·g−1, respectively. These molybdenum oxides weak electrical conductivity (MoO3 = 9.6 × 10−7 s·cm−1, MoO2 = 8.8 × 10−5 s·cm−1) and slow diffusion kinetics of Li+ ions, however, are often their limitations [24]. In addition, the molybdenum oxides undergo a significant volume expansion of around 100–250% during insertion and extraction. This causes internal tensions, electrode pulverization, and loss of interparticle contact, which leads to electrical disconnection [24,63,75]. Poor rate performance, severe capacity fading, and suboptimal cycling stability are the results of these problems [63,75]. However, there has been an increase in the past 20 years in papers specifically about MoO3 and MoO3-based systems used for lithium-ion batteries (Figure 9), with many studies concentrating on solving these difficulties because of the benefits of these materials. Table 2 lists recent studies on MoO3 and MoO3-based systems for LIBs.
To get around the problems with molybdenum oxides and get stable cycling with a higher reversible capacity and rate performance, researchers have looked into a number of effective strategies, especially to lower the diffusion barrier and band gap and shorten the electron and Li+ diffusion channels [76]. When nanostructures like nanosheets, nanoparticles, nanobelts, and nanowires are made, it is easier for electrolyte ions to reach the active sites [75,76]. On the other hand, bandgap and diffusion barriers have been lowered via vacancy engineering and heteroatom doping. The creation of oxygen vacancies has been suggested as a way to improve lithium storage efficiency since they can dramatically boost ion diffusion kinetics, increase electrical conductivity, and provide more active sites for redox reactions [76]. Lithium ions can be inserted and removed during charging and discharging cycles at the vacant sites in the MoO3 crystal intralayer and interlayers [68].
Free electrons present in the MoO3 conduction bands are primarily responsible for its electrical conductivity, and the intensities of these free electrons in such material are significantly influenced by the concentration of defects, such as oxygen deficiency [1]. By electron-proton co-doping between low-work-function metals and MoO3 in a deionized water environment, it is possible to create oxygen vacancies. The hydrogen doping preferentially chooses symmetric oxygen to generate the unstable OH group, which distorts the MoO3 lattice. By releasing this unstable OH group as H2O into the solution, MoO3 loses oxygen from its lattice and forms oxygen vacancies [76]. Two-dimensional (2D) MoO3 materials have also attracted great interest due to the shortened lithium-ion transportation path and released stress concentration. These materials also provide more sites for lithium insertion and can significantly increase lithium storage capacity [75].
In addition, MoO3 nanostructured materials engineered by carbon coatings would surely bring advantages for electrochemical performances due to the potential capability of the synergetic effects of each component [24]. Carbonaceous materials such as graphene, amorphous carbon, carbon nanotube (CNT), and graphite can improve structural stability, compensate for the low electrical conductivity of MoO3 anodes, and accommodate the large strain induced by Li-ion diffusion during cycling [62]. Conversion reactions are involved in the lithiation/delithiation mechanism for the layered MoO3 [24,63]:
M o O 3 + x L i + + x e L i x M o O 3
L i x M o O 3 + 6 x L i + + 6 x e 3 L i 2 O + M o
The rutile-type structure of MoO2 features a high local defect concentration that changes the atomic arrangement, facilitating the transport and storage of Li+ ions. Unlike MoO3 and similar to the graphite mechanism, MoO2 follows just a one-step reversible reaction [24]:
M o O 2 + x L i + + x e L i x M o O 2
Table 2. MoO3 and MoO3-based systems used in LIBs applications.
Table 2. MoO3 and MoO3-based systems used in LIBs applications.
MaterialMorphologySynthesis MethodCurrent Density (mA·g−1)Li Ion Capacity (mA·h.g−1) aCyclesInitial Coulombic Efficiency (ICE) (%)Reference
α-MoO3SheetsThermal plasma300070020070[75]
α-MoO3NanobeltsHydrothermal10078710070.6[77]
α-MoO3−xNanobeltsMechanical grinding500820200UD b[78]
N-α-MoO3−xNanoflowersNR c100126145081.2[76]
Carbon-α-MoO3Nanoparticles/NanofibersElectrospinning500801.1200~98[4]
Carbon-α-MoO3-MoO2Nanoribbons/NanoparticlesSolid-phase10084010052[24]
CNT-α-MoO3Nanofibers/NanoplatesElectrospinning100097210062.7[62]
α-MoO3-SnS2Nanorods/NanosheetsHydrothermal60568.210092.7[79]
Carbon-MoO2Nanoplate-likeCO2 oxidation50323300UD b[80]
α-MoO3-TiO2NanobeltsHydrothermal400935.840073[81]
α-MoO3-NiOFlower-like microspheresHydrothermal + Coverage10094410081[82]
Carbon-α-MoO3-SnO2NanoflakesChemical Vapor Deposition + Dispersion2001020.520062.8[83]
α-MoO3-RGONanoparticlesUltrasonication500568100UD b[84]
CNT-SiO2-α-MoO3Cactus-likeSelf-assembly and in situ carbonization100070050099.8[85]
Fe2O3-TeO2-MoO3NanoparticlesMolten1000463.280047.4[86]
Carbon-SnO2-MoO3Nanoparticles/NanosheetsHydrothermal2001338.330070[87]
h-MoO3-GrafeneMicrorodsScalable precipitation200066530063[88]
h-MoO3-GOMicrorodsScalable precipitation100078910065[88]
SnO2-MoO3-GrapheneNano-grain/Sheet-likeHydrothermal2001522.525072.2[89]
SnO2-MoO3-CNTNanoparticles/NanotubesHydrothermal2001372.228080.9[90]
Carbon-α-MoO3NanoparticlesSolution combustion100668200UD b[91]
α-MoO3-Fe2O3Micro-octahedronsThermolysis 200121835081.9[92]
α-MoO3-MoO2-g-C3N4NanosheetsHydrothermal + Freeze-Drying50099210072[93]
a Value after the cycling test. b UD: Unavailable Data. c NR: Numerical Approach.

4. Water Pollution Treatment: Adsorption and Photocatalysis

Global industrial development has led to undesirable environmental problems, and one of the worst is water pollution. Among the water pollutants, heavy metal ions are harmful contaminants because of their high toxicity and mutagenicity/carcinogenicity potential, even at very low concentrations [94,95]. Organic toxic substances and several heavy metals are discharged into the environment without appropriate treatment. These pollutants originate from the paper, textile, food, pharmaceutical, printed circuit board manufacturing, and solar cell industries [15,96,97]. Because of this, a high number of studies have focused on the development of more efficient, low-cost, easy-to-operate, environmentally friendly, and reusable materials for water treatment. The adsorbents and the adsorption process have become two of the principal methods to remediate this problem. However, the search for new, low-cost, and more effective sorbents is still a challenge worldwide [95,97,98].
Molybdenum-based materials have been indicated as efficient materials for wastewater treatment due to their attractive and unique adsorptive, catalytic, optical, and electronic properties, in addition to their high mechanical, thermal, and chemical stability [97]. According to Figure 10, an increasing trend in the last 20 years in papers directly related to MoO3 and MoO3-based systems applied to adsorption has been verified.
Oxygen vacancies are often formed on the surface with a negative zeta potential, and consequently, strong electrostatic adsorption between MoO3−x and cationic dyes is observed [98,99]. In Table 3, recent studies on MoO3 and MoO3-based systems in the adsorption process are shown. Interestingly, these materials present high adsorptive capacity, especially towards cationic dyes such as methylene blue (MB) [22,97,98,100,101,102,103], crystal violet (CV) [101], and malachite green (MG) [101]. In addition, heavy metal ions such as Pb2+ [94,95,97,104], Cr3+ [102], Cu2+ [104], Mn2+ [105], Cr6+ [105], and aromatic sulfur compounds like thiophene [106] and dibenzothiophene [106] are removed by the electrostatic adsorption interactions. Removal efficiency of some metal ions with an a-MoO3 nanosheet array system is studied by Yunying Wu et al. [94], showing a higher adsorption capacity toward Pb2+ from an aqueous solution than Cu2+, Zn2+, Cr3+, and Cd2+ (see Figure 11). The enhanced performance in Pb2+ removal can be attributed to the electrostatic adsorption interactions between oxygen-containing functional groups and Pb2+, besides the partial formation of coordination bonds between oxygen groups and Pb2+ and also the formation of a new substance, the PbMoO4 precipitate [94,97], from a solid-liquid interfacial reaction as shown in Equation (4) [94]:
M o O 3 + H 2 O + P b 2 + P b M o O 4 + 2 H +
Kedves et al. observed that the α-MoO3 adsorption capacity has a direct relationship with the cationic group quantity [101]. The presence of imine, two amino, and thiol groups in the molecule of methylene blue favored faster adsorption by the α-MoO3-based system. On the contrary, crystal violet with three amine functional groups, malachite green with only two, and rhodamine B with two amine groups with a carboxylic and an ether group added resulted in reduced adsorption efficiency. Thus, the adsorption effectiveness declines as the number of cationic groups decreases. The presence of more positively charged functional groups in methylene blue tends towards greater bond formation with the negatively charged α-MoO3 surface.
For anionic dye adsorption, an interesting strategy is to produce composites by intercalating the favorable layered structure of MoO3 with active sites charged positively. Wang et al. [107] prepared the Al13-3.34%@MoO3 composite by intercalating polycationic Al13 into an extensive airspace of MoO3. The anionic dye methyl orange (MO) was rapidly adsorbed due to the strong electrostatic interaction between the cation Al13 and the anion MO.
MoO3 materials are also effective in removing metallic ions such as Pb2+. The electrostatic interaction between O-containing functional groups and Pb2+, as well as the partial creation of coordination bonds between oxygen groups and Pb+2, govern the process of Pb2+ removal by MoO3. The thermodynamic parameters [97] suggest that the Pb2+ removal on α-MoO3 is spontaneous and endothermic, and the Pb+2 removal matches quite well with the Langmuir isotherm. The pH solution interferes with the Pb+2 adsorption process. In the range (2.0–8.0), there are three forms of lead species: Pb2+ (dominant at pH < 6.0), Pb(OH)+ (formed from the Pb2+ hydrolysis started at pH = 3.7 and dominant at pH > 7.5), and Pb(OH)2 (formed from the Pb2+ hydrolysis started at pH = 6.8) [108]. The Pb2+ removal slightly increased upon increasing the pH from 2.0 to 5.0, and the removing rate reached its highest value at pH = 5.0. However, it slightly decreased when the pH increased from 7.0 to 8.0. The results suggested that the electrostatic interactions occurred between α-MoO3 and Pb2+/Pb(OH)+ ions. Thus, α-MoO3 exhibited a high removal capacity over a broad pH range due to its negative surface charge in this pH interval, presenting high selectivity toward Pb2+ compared to Cu2+, Zn2+, Cr3+, and Cd2+ ions [94].
Molybdenum trioxide (MoO3) is an n-type semiconductor that has been used as an active photocatalyst because of its effective absorption of the majority of UV light and partial visible light absorption, with a wide optical band gap (2.8–3.6 eV) that varies depending on the growth conditions and synthesis method [18,72,109]. An increasing trend in the last 20 years in papers directly related to MoO3 and MoO3-based systems applied to photocatalysis is verified in Figure 12. Recent studies on MoO3 and MoO3-based systems used in photocatalytic applications are summarized in Table 4. Photocatalytic applications benefit from the multiple Mo valence states, chemical stability, non-toxicity, ability to transport holes, and high aqueous medium stability [110]. In particular, MoO3 materials have a strong oxidation capability in the photocatalytic process due to their low valence band level [19]. However, MoO3 presents limited photocatalytic effects due to its fast electron-hole recombination and limited visible light absorption [111,112,113]. To solve these issues, these semiconductors are doped, which is considered a promising strategy to improve photocatalytic performance [111,114]. The heterojunction formed by the different individual band gaps creates intermediate energy levels in the gap region, which facilitate the electronic transition and provide efficient separation of electron-hole pairs and recombination retardation [112]. This occurs through a rectifying effect at the heterojunction interface and promotes the enhancement of the photocatalytic reactions. The heterostructures also provide large interfacial areas for the occurrence of redox reactions [112]. Thereby, heterostructured systems doped with Fe2O3 [15,111], rGO [15], SnO2 [115], Cu2S [116], WO3 [110,115], CdTiO3 [117], SiO2 [118], g-C3N4 [58,112,119], NiO [120], polypyrene [121], etc., and also metal doping with Ag [118] were synthesized in the last years for photocatalytic applications such as dye and antibiotic degradation and heavy metal reduction (Cr6+), showing excellent performance for RhB [58,112,118], MB [15,115,116,120,122], tetracycline [111], acetaminophen [121], and metronidazole photodegradation under visible light irradiation [118], besides aspirin degradation under UV irradiation [117] and MB and 2,4-dichlorophenol under sunlight irradiation [119].
The electronic band of the MoO3 and crystal structure have an important effect on photocatalytic performance [19]. Thus, oxygen vacancies are thought to be an important contributor to this process’s effectiveness. These vacancies function as an electron trap, capturing the photogenerated electrons and suppressing the photoexcited carrier recombination [58,123,124]. “Vacancy doping” consists of the strategy of introducing free holes or free electrons for a higher charge carrier density without the need for extrinsic dopant ions [123]. MoO3 materials allow an adjustable oxygen vacancy density due to the special nature of their outer-d valence electrons, and these oxygen vacancies in MoO3−x cause an extra defect band below the conduction band that decreases its effective bandgap as a consequence of the creation of the intermediate band (IB) in the gap [1,23,124]. The insertion of oxygen vacancies causes the creation of Z-type heterojunctions, and MoO3−x-based materials can generate high-energy electrons to participate in redox reactions [58]. Some synthetic methods for inserting oxygen vacancies include chemical reduction, ion intercalation, chemical vapor deposition, and ball milling treatment [1,124,125]. By modifying and altering the ball-milling periods, for instance, the oxygen vacancy content on the surface of α-MoO3 can be adjusted [125]. In the synthesis of MoO3−x, oleylamine is also reported to be utilized as a weak reducing agent, causing the formation of Mo5+ and consequent oxygen vacancies [19].
Different heterostructures can be synthesized with MoO3−x for higher photocatalytic performances. For example, the matching band structure of sulfur-doped carbon nitride (S-CN) and MoO3−x is expected to construct an S-scheme heterojunction catalyst. Because S and O atoms have similar electronegativity, during heterojunction creation, lone pair electrons of S atoms on S-CN can occupy the oxygen vacancy electron defect state on the MoO3−x surface. By regulating the S-CN ratio, it is possible to maintain the proper oxygen vacancies while controlling the surface defect density of MoO3−x. This can help promote charge carrier separation as well as the adsorption and activation of molecular oxygen. The metal Mo atoms in MoO3−x can interact with the lone pair electrons of the N atoms in S-CN, acting as electron donors, to form a covalent Mo(δ+)–N(δ-) bond with minimal space constraints. The charge transfer barrier can be lowered and the charge transfer rate accelerated by using this interfacial chemical bond as a particular “highway” [126].
In addition, researchers have been interested in improving the photocatalytic activity of polymeric carbon nitride (PCN) by coupling it with molybdenum-based compounds employed as co-catalysts [123]. Because g-C3N4 is positively charged and MoO3 is negatively charged, an electric field is formed between them. This built-in electric field, along with the bending direction of the band edge because of the nature of the electrons obtained, results in a Z-scheme carrier transfer mechanism. This indicates that the photogenerated electrons in the MoO3 conduction band can easily move and recombine with the holes in the g-C3N4 valence band under light irradiation, facilitating photogenerated carrier separation efficiency [112].
Embedded Mo reduces the PCN band gap and increases its valence band position, which improves visible light absorption and oxidation ability. When exposed to visible light irradiation, the Mo-embedded PCN creates photogenerated carriers, and the embedded Mo6+, as an electron transfer carrier, reduces the recombination. Three factors improve the photocatalytic activity considerably. First, the embedded Mo6+ reduces the recombination of photogenerated carriers by shortening the migration distance of electrons from the excited to the reactive sites. Second, the embedded Mo6+ lowers the PCN band gap, allowing it to capture more visible light irradiation. The valence band potential of PCN is increased, and the oxidizing ability of the photogenerated holes is improved. Third, more adsorptive and photocatalytic active sites emerge from the increased specific surface area. Because of the suitable and tunable band gaps between the semiconductors, their larger surface area, platinum-like electronic structure, and high electron storage capacity, semiconductors containing Mo, such as MoO3 and MoO2, can improve the interfacial transport of photogenerated carriers, resulting in improved photocatalytic activity. The chemistry related to the multivalent nature of Mo can explain the Mo effect. The electron mobility of the system is improved, and the internal resistance is reduced because of the Mo species with abundant d-electrons [127].
Table 4. MoO3 and MoO3-based systems used in photocatalysis applications.
Table 4. MoO3 and MoO3-based systems used in photocatalysis applications.
MaterialMorphologySynthesis Method Adsorbent NatureTargetLamp/PowerTime (min)Removal Efficiency (%)Reference
α-MoO3NanoparticlesHydrothermalCationic dyeMBLED (12 W)9099[128]
α-MoO3Plate-likePechini-basedCationic dyeRhB160 W1593.3[129]
α-MoO3NanorodsHydrothermalCationic dyeMBXenon (300 W)12098[130]
α-MoO3NanorodsHydrothermalCationic dyeRhBXenon (300 W)12090[130]
α-MoO3MicrofibersHydrothermalCationic dyeMBHalogen (150 W)12090[131]
g-C3N4-tourmaline-α-MoO3−xNanosheetsHydrothermalCationic dyeRhBXenon (300 W)40~100[58]
g-C3N4-α-MoO3Nanosheets/NanoparticlesMixing/AnnealingCationic dyeRhBXenon (300 W)24100[112]
α-MoO3-Cu2SNanoparticles/Hexagonal rodsHydrothermalCationic dyeMBMetal-halide
(200 W)
12098[116]
α-MoO3-SiO2MonolithsSol–Gel/Wet impregnationCationic dyeRhBCFL/(45 W)15088.6[118]
α-MoO3-NiONanobelts/NanoparticlesHydrothermalCationic dyeMBXenon
(500 W)
12096.5[120]
α-MoO3NanoparticlesHydrothermalAnionic dyeEosin YellowLED (12 W)9094[128]
α-MoO3NanorodsHydrothermalAnionic DyeAlizarinXenon (300 W)12074[130]
Ag-α-MoO3-TiO2Spherical particlesHydrothermalAnionic dyeMOUV light (100 W)33075.8[132]
g-C3N4-α-MoO3Nanosheets/NanoparticlesMixing/AnnealingAnionic dyeMOXenon (300 W)4082.3[112]
α-MoO3-WO3NanorodsHydrothermalMetal ionCr6+Tungsten
(150 mW/cm−2)
2597.6[110]
α-MoO3@ZIF-8Nanowires/nanoparticlesHydrothermalMetal ionCr6+Sun 200040100[133]
MWCNT-α-MoO3Nanotubes/NanoparticlesHydrothermalMetal ionCr6+Xenon (300 W)30100[134]
α-MoO3-Fe2O3NanorodsHydrothermalDrugTetracyclineUD a9096.5[111]
α-MoO3-SiO2MonolithsSol-Gel/Wet impregnationDrugMetronidazoleCFL/(45 W)18067.4[118]
PPy-α-MoO3Spherical like
Particles/Platelike
In situ polymerizationDrugAcetaminophenXenon
(500 W)
18093.4[121]
g-C3N4-α-MoO3ParticlesImpregnation and calcinationHerbicide2,4-dichlorophenolNatural sunlight30099[119]
a Unavailable data.
Some organic dye photodegradation mechanisms over MoO3 materials can be proposed. For example, the MB photodegradation possibly occurs through α-MoO3 photoexcitation, resulting in an electron-hole pair generation on the surface [131]. Through the use of the corresponding trapping scavenges, it is indicated that hole radicals (h+) and hydroxyl radicals (OH) are the primary reactive species acting as strong oxidants. On the other hand, according to the literature, the superoxide anion radicals ( O 2 . ) have a minor role in this photodegradation process [116,131]. The redox reactions that take place in the visible-light photocatalytic mechanism for MB photodegradation are schematically represented in Figure 13. In the RhB photodegradation, the superoxide anion radicals ( O 2 . ) and hole radicals (h+) play the main role as active species, with a small number of hydroxyl groups (OH) also involved in the process reactions [111,130]. Thus, only some of the holes are used to oxidize water, while most of them are used for RhB photodegradation [58]. The phenomenon obeys an indirect mechanism, with strong oxidizing agents generated on the surface of the catalyst. These oxidizing agents react homogeneously with the organic material, resulting in its degradation [128]. During the Alizarin (AZ) photodegradation, the superoxide radicals ( O 2 . ) act as active radical species, while the hole radicals (h+) do not act as active radicals [130].

5. Conclusions

This brief review comprised the latest advances in technological applications of MoO3 and MoO3-based materials in gas sensors, lithium-ion batteries, adsorption, and photocatalysis. Due to its electrochemical activity, high dielectric constant, great gas molecules adsorption, and charge carrier concentration modulation in response to gases, MoO3 is strongly indicated as a potential material as a gas sensor for NO2, NO, H2, CO, C2H5OH, NH3, trimethylamine, triethylamine, acetone, 1-butylamine, and water (humidity sensor), presenting a different sensing mechanism in comparison with typical n-type semiconductors. The reactions proposed in the MoO3 materials are a lattice oxygen reaction instead of a traditional surface-chemisorbed oxygen mechanism due to their crystalline-layered structure. However, the insufficient electrical conductivity of MoO3 restricts its electrochemical performance, stimulating investigations aimed at improving its performance with functional and surface modifications, morphological and sizing adjustments, the introduction of doping elements, the loading of noble metals, and the synthesis of composites. The MoO3 crystal is advantageous for lithium-ion intercalation because of the weak Van Der Waals force between the stacked sheets along the [0 1 0] direction. The intercalation ability and layered structure also provide diffusing channels and embedding sites for Li+ ions, making molybdenum oxides excellent options as cathode and anode materials for lithium-ion batteries, presenting higher theoretical specific capacities and faster discharge potential compared to conventional graphite. In this context, the weak electrical conductivity and slow diffusion kinetics of Li+ ions are often the limitations of MoO3. However, to overcome these problems, effective strategies, such as the diffusion barrier and band gap reduction, and the shortening of the electron and Li+ diffusion channels, have been researched. Molybdenum-based materials have also been reported for efficient photocatalytic applications, showing great promise in dye and antibiotic degradation under visible and ultraviolet irradiation. In particular, the oxygen vacancy creation induces an additional defect band that decreases its effective bandgap as a consequence of the intermediate band creation in the gap, acting as an electron trap and retarding the recombination. In particular, these materials have a strong oxidation capacity in the photocatalytic process due to their low valence band level. However, their rapid electron-hole recombination and limited absorption of visible light impose restrictions on their applications. To overcome these challenges and improve photocatalytic performance, doping of these semiconductors is a considered strategy since the formation of heterojunctions between different individual band gaps creates intermediate energy levels in the band gap region, facilitating the electronic transition and promoting efficient separation. of electron-hole pairs, in addition to delaying recombination. MoO3-based materials also present high adsorption performance and efficiency for pollutants removal through adsorption processes due to the abundant and fully exposed active sites on the oxygen-rich structure of MoO3, presenting high adsorptive capacity, especially towards cationic dyes because of its negatively charged surface, and heavy metal ions such as Pb2+, Cr3+, Cr6+, Cu2+, and Mn2+. On the other hand, for the adsorption of anionic dyes, research has revealed a viable option to produce composites by interspersing the layered structure of MoO3 with positively charged active sites. It is then possible to confirm the high applicability of MoO3-based materials in recent and important technological applications. We mainly outlined the applications of MoO3 in the fields of energy conversion devices, gas sensors, and catalysts. With regard to energy storage and conversion devices, more promising studies must be conducted since they have a bottleneck to ensure both capacitance and durability. As for gas sensors, though the MoO3 exhibits a good response to various gases, it is still necessary to investigate the sensing performance under different humid conditions to establish the relationship between the sensing properties and the environmental conditions. The catalytic nature and mechanism of MoO3-based catalysts are not fully understood, and accurate studies on the structure, phase, morphology, hybridization, and oxygen vacancies of MoO3-based materials must be conducted.

Author Contributions

Conceptualization, M.G.d.S.J., L.C.C.A., H.B.S., R.M.d.C.F. and R.R.M.; formal analysis and writing, M.G.d.S.J., L.C.C.A. and R.R.M.; review and supervision, R.R.M., G.d.A.N. and H.d.L.L.; project administration, funding acquisition, R.R.M., G.d.A.N. and H.d.L.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the Brazilian funding agencies CNPq (National Council for Scientific and Technological Development—grants 420004/2018-1, 309771/2021-8, and 303680/2019-9) and CAPES.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study, in the collection, analyses, or interpretation of data, in the writing of the manuscript, or in the decision to publish the results.

References

  1. de Castro, I.A.; Datta, R.S.; Ou, J.Z.; Castellanos-Gomez, A.; Sriram, S.; Daeneke, T.; Kalantar-Zadeh, K. Molybdenum Oxides—From Fundamentals to Functionality. Adv. Mater. 2017, 29, 1701619–1701649. [Google Scholar] [CrossRef] [PubMed]
  2. Alizadeh, S.; Hassanzadeh-Tabrizi, S.A. MoO3 fibers and belts: Molten salt synthesis, characterization and optical properties. Ceram. Int. 2015, 41, 10839–10843. [Google Scholar] [CrossRef]
  3. Kim, M.R.; Ahn, S.J.; Jang, D.J. Fabrication and characterization of MoO3 clusters-coated TiO2 nanotubes showing charge-transferred photoluminescence. Eur. Phys. J. D 2009, 52, 75–78. [Google Scholar] [CrossRef]
  4. Zhao, X.; Jia, W.; Wu, X.; Lv, Y.; Qiu, J.; Guo, J.; Wang, X.; Jia, D.; Yan, J.; Wu, D. Ultrafine MoO3 anchored in coal-based carbon nanofibers as anode for advanced lithium-ion batteries. Carbon 2020, 156, 445–452. [Google Scholar] [CrossRef]
  5. Nadimicherla, R.; Zha, R.; Wei, L.; Guo, X. Single crystalline flowerlike α-MoO3 nanorods and their application as anode material for lithium-ion batteries. J. Alloys Compd. 2016, 687, 79–86. [Google Scholar] [CrossRef]
  6. Singh, A.; Kumar, S.; Bawankule, P.; Gupta, A.; Kumar, R.; Kumar, P.; Tiwari, A. Selective wavelength optical filters from mixed polymorph and binary integration of MoO3 multilayer structures. Opt. Mater. 2021, 111, 110709–110718. [Google Scholar] [CrossRef]
  7. Ijeh, R.O.; Nwanya, A.C.; Nkele, A.C.; Madiba, I.G.; Bashir, A.K.H.; Ekwealor, A.B.C.; Osuji, R.U.; Maaza, M.; Ezema, F. Optical, electrical and magnetic properties of copper doped electrodeposited MoO3 thin films. Ceram. Int. 2020, 46, 10820–10828. [Google Scholar] [CrossRef]
  8. Koçak, S.; Ertaş, F.N.; Dursun, Z. Electrochemical deposition and behavior of mixed-valent molybdenum oxide film at glassy carbon and ITO electrodes. Appl. Surf. Sci. 2013, 265, 205–213. [Google Scholar] [CrossRef]
  9. Soultati, A.; Kostis, I.; Argitis, P.; Dimotikali, D.; Kennou, S.; Gardelis, S.; Speliotis, T.; Kontos, A.G.; Davazoglou, D.; Vasilopoulou, M. Dehydration of molybdenum oxide hole extraction layers via microwave annealing for the improvement of efficiency and lifetime in organic solar cells. J. Mater. Chem. C 2016, 4, 7683–7694. [Google Scholar] [CrossRef]
  10. Vasilopoulou, M.; Soultati, A.; Argitis, P.; Stergiopoulos, T.; Davazoglou, D. Fast recovery of the high work function of tungsten and molybdenum oxides via microwave exposure for efficient organic photovoltaics. J. Phys. Chem. Lett. 2014, 5, 1871–1879. [Google Scholar] [CrossRef]
  11. Vasilopoulou, M.; Douvas, A.M.; Georgiadou, D.G.; Palilis, L.C.; Kennou, S.; Sygellou, L.; Soultati, A.; Kostis, I.; Papadimitropoulos, G.; Davazoglou, D.; et al. The influence of hydrogenation and oxygen vacancies on molybdenum oxides work function and gap states for application in organic optoelectronics. J. Am. Chem. Soc. 2012, 134, 16178–16187. [Google Scholar] [CrossRef]
  12. Chithambararaj, A.; Bose, A.C. Hydrothermal synthesis of hexagonal and orthorhombic MoO3 nanoparticles. J. Alloys Compd. 2011, 509, 8105–8110. [Google Scholar] [CrossRef]
  13. Kolawole, F.O.; Ramirez, M.A.; Kolawole, S.K.; Varela, L.B.; Tschiptschin, A.P. Deposition and characterization of molybdenum oxide (MoO3) nanoparticles incorporated diamond-like carbon coatings using pulsed-DC PECVD. Mater. Lett. 2020, 278, 128420. [Google Scholar] [CrossRef]
  14. Lemos, R.M.J.; Balboni, R.D.C.; Cholant, C.M.; Azevedo, C.F.; Pawlicka, A.; Gündel, A.; Flores, W.H.; Avellaneda, C.O. Molybdenum doping effect on sol-gel Nb2O5: Li+ thin films: Investigation of structural, optical and electrochromic properties. Mater. Sci. Semicond. Process. 2021, 134, 105995. [Google Scholar] [CrossRef]
  15. Anjaneyulu, R.B.; Mohan, B.S.; Naidu, G.P.; Muralikrishna, R. Visible light enhanced photocatalytic degradation of methylene blue by ternary nanocomposite, MoO3/Fe2O3/rGO. J. Asian Ceram. Soc. 2018, 6, 183–195. [Google Scholar] [CrossRef]
  16. Sreedhara, M.B.; Matte, H.S.; Govindaraj, A.; Rao, C.N. Synthesis, characterization, and properties of few-layer MoO3. Chem. Asian J. 2013, 8, 2430–2435. [Google Scholar] [CrossRef] [PubMed]
  17. Koike, K.; Wada, R.; Yagi, S.; Harada, Y.; Sasa, S.; Yano, M. Characteristics of MoO3 films grown by molecular beam epitaxy. Jpn. J. Appl. Phys. 2014, 53, 05FJ02. [Google Scholar] [CrossRef]
  18. Han, Y.; Rheem, Y.; Lee, K.-H.; Kim, H.; Myung, N.V. Synthesis and characterization of orthorhombic-MoO3 nanofibers with controlled morphology and diameter. J. Ind. Eng. Chem. 2018, 65, 231–238. [Google Scholar] [CrossRef]
  19. Huang, Z.; Liu, J.; Zong, S.; Wang, X.; Chen, K.; Liu, L.; Fang, Y. Fabrication of graphitic carbon Nitride/Nonstoichiometric molybdenum oxide nanorod composite with the nonmetal plasma enhanced photocatalytic hydrogen evolution activity. J. Colloid Interface Sci. 2022, 606, 848–859. [Google Scholar] [CrossRef]
  20. Di Yao, D.; Field, M.R.; O’Mullane, A.P.; Kalantar-Zadeh, K.; Ou, J.Z. Electrochromic properties of TiO2 nanotubes coated with electrodeposited MoO3. Nanoscale 2013, 5, 10353–10359. [Google Scholar] [CrossRef]
  21. Samsuri, A.; Latif, M.N.; Shamsuddin, M.R.; Salleh, F.; Abu Tahari, M.N.; Tengku Saharuddin, T.S.; Dzakaria, N.; Yarmo, M.A. Studies on influence of hydrogen and carbon monoxide concentration on reduction progression behavior of molybdenum oxide catalyst. Int. J. Hydrogen Energy 2021, 46, 24831–24844. [Google Scholar] [CrossRef]
  22. Taghipour, E.; Tahmasebi, N. One-step hydrothermal synthesis of MoO3/MoO2 composite for efficient removal of methylene blue from aqueous solutions. Mater. Today Commun. 2021, 26, 102012–102020. [Google Scholar] [CrossRef]
  23. Chen, C.Z.; Li, Y.; Tang, X.D. Evidence of oxygen vacancy and possible intermediate gap state in layered α-MoO3 single-crystal nanobelts. Phys. B Condens. 2016, 481, 192–196. [Google Scholar] [CrossRef]
  24. Borgschulte, A.; Sambalova, O.; Delmelle, R.; Jenatsch, S.; Hany, R.; Nüesch, F. Hydrogen reduction of molybdenum oxide at room temperature. Sci. Rep. 2017, 7, 40761–40769. [Google Scholar] [CrossRef] [PubMed]
  25. Rodríguez-Gattorno, G.; Martínez-Hernández, A.; Aleman-Vázquez, L.O.; Torres-García, E. Structural and thermal study of carbon-modified molybdenum sub-oxide catalysts. Appl. Catal. A Gen. 2007, 321, 117–124. [Google Scholar] [CrossRef]
  26. Dang, J.; Zhang, G.-H.; Chou, K.-C.; Reddy, R.G.; He, Y.; Sun, Y. Kinetics and mechanism of hydrogen reduction of MoO3 to MoO2. Int. J. Refract. Met. Hard Mater. 2013, 41, 216–223. [Google Scholar] [CrossRef]
  27. Wang, L.; Zhang, G.-H.; Chou, K.-C. Mechanism and kinetic study of hydrogen reduction of ultra-fine spherical MoO3 to MoO2. Int. J. Refract. Hard Met. 2016, 54, 342–350. [Google Scholar] [CrossRef]
  28. Zhu, Y.; Yao, Y.; Luo, Z.; Pan, C.; Yang, J.; Fang, Y.; Deng, H.; Liu, C.; Tan, Q.; Liu, F.; et al. Nanostructured MoO3 for efficient energy and environmental catalysis. Molecules 2020, 25, 18. [Google Scholar] [CrossRef]
  29. Ren, H.; Sun, S.; Cui, J.; Li, X. Synthesis, functional modifications, and diversified applications of molybdenum oxides micro-/nanocrystals: A review Cryst. Growth Des. 2018, 18, 6326–6369. [Google Scholar] [CrossRef]
  30. Malik, R.; Joshi, N.; Tomer, V.K. Advances in the designs and mechanisms of MoO3 nanostructures for gas sensors: A holistic review. Mater. Adv. 2021, 2, 4190–4227. [Google Scholar] [CrossRef]
  31. Bisht, P.; Kumar, A.; Jensen, I.T.; Ahmad, M.; Belle, B.D.; Mehta, B.R. Enhanced gas sensing response for 2D α-MoO3 layers: Thickness-dependent changes in defect concentration, surface oxygen adsorption, and metal-metal oxide contact. Sens. Actuators B Chem. 2021, 341, 129953–129963. [Google Scholar] [CrossRef]
  32. Wang, J.; Zhou, Q.; Peng, S.; Xu, L.; Zeng, W. Volatile Organic Compounds Gas Sensors Based on Molybdenum Oxides: A Mini Review. Front. Chem. 2020, 8, 339–345. [Google Scholar] [CrossRef] [PubMed]
  33. Bhat, S.; John, A.; Panda, P.K. A review on metal-oxide based p-n and n-n heterostructure nanomaterials for gas sensing applications. Sens. Int. 2021, 2, 100085–100103. [Google Scholar]
  34. Boboriko, N.E.; Dzichenka, Y.U. Molecular dynamics simulation as a tool for prediction of the properties of TiO2 and TiO2:MoO3 based chemical gas sensors. J. Alloys Compd. 2021, 855, 157490–157501. [Google Scholar] [CrossRef]
  35. Mo, Y.; Tan, Z.; Sun, L.; Lu, Y.; Liu, X. Ethanol-sensing properties of α-MoO3 nanobelts synthesized by hydrothermal method. J. Alloys Compd. 2020, 812, 152166–152170. [Google Scholar] [CrossRef]
  36. Farzi-kahkesh, S.; Fattah, A.; Rahmani, M.B. Synthesis and optimum temperature determination of highly sensitive MoO3-based heterojunction Schottky sensor for hydrogen detection. Microelectron. Eng. 2021, 235, 111453–111459. [Google Scholar] [CrossRef]
  37. Fu, H.; Wu, Z.; Yang, X.; He, P.; An, X.; Xiong, S.; Han, D. Ultra-high sensitivity and selectivity of Au nanoparticles modified MoO3 nanobelts towards 1-butylamine. Appl. Surf. Sci. 2021, 542, 148721–148731. [Google Scholar] [CrossRef]
  38. Li, W.; He, S.; Feng, L.; Yang, W. Cr-doped α-MoO3 nanorods for the fast detection of triethylamine using a pulse-heating strategy. Mater. Lett. 2019, 250, 143–146. [Google Scholar] [CrossRef]
  39. Sakhuja, N.; Jha, R.; Bhat, N. Facile green synthesis of 2D hexagonal MoO3 for selective detection of ammonia at room temperature. Mater. Sci. Eng. B. 2021, 271, 115249–115259. [Google Scholar] [CrossRef]
  40. Li, X.; Yang, H.; Hu, X.; Wu, Q.; Xiong, W.; Qin, Z.; Xie, C.; Zeng, D. Exposed Mo atoms induced by micropores enhanced H2S sensing of MoO3 nanoflowers. J. Hazard. Mater. 2022, 429, 128270–128278. [Google Scholar] [CrossRef]
  41. Bai, S.; Chen, S.; Chen, L.; Zhang, K.; Luo, R.; Li, D.; Liu, C.C. Ultrasonic synthesis of MoO3 nanorods and their gas sensing properties. Sens. Actuators B Chem. 2012, 174, 51–58. [Google Scholar] [CrossRef]
  42. Li, W.; Ou, Q.; Wang, X.; Xing, K.; Tesfamichael, T.; Motta, N.; Qi, D.-C. Large-sized α-MoO3 layered single crystals for superior NO2 gas sensing. Appl. Surf. Sci. 2022, 586, 152793–152804. [Google Scholar] [CrossRef]
  43. Wang, J.; Zhou, Q.; Wei, Z.; Xu, L.; Zeng, W. Experimental and theoretical studies of Zn-doped MoO3 hierarchical microflower with excellent sensing performances to carbon monoxide. Ceram. Int. 2020, 46, 29222–29232. [Google Scholar] [CrossRef]
  44. Thomas, T.; Kumar, Y.; Ramos Ramón, J.A.; Agarwal, V.; Guzmán, S.S.; Reshmi, R.; Pushpan, S.; Loredo, S.L.; Sanal, K.C. Porous silicon/α-MoO3 nanohybrid based fast and highly sensitive CO2 gas sensors. Vacuum 2021, 184, 109983–109993. [Google Scholar] [CrossRef]
  45. Ji, H.; Zeng, W.; Li, Y. New insight into the gas-sensing properties of nanofiber-assembled and nanosheet-assembled hierarchical MoO3 structures. Phys. E Low-Dimens. Syst. Nanostruct. 2019, 114, 113646–113652. [Google Scholar] [CrossRef]
  46. Li, J.; Liu, H.; Fu, H.; Xu, L.; Jin, H.; Zhang, X.; Wang, L.; Yu, K. Synthesis of 1D α-MoO3/0D ZnO heterostructure nanobelts with enhanced gas sensing properties. J. Alloys Compd. 2019, 788, 248–256. [Google Scholar] [CrossRef]
  47. Thomas, T.; Jayababu, N.; Shruthi, J.; Mathew, A.; Cerdán-Pasarán, A.; Hernández-Magallanes, J.A.; Sanal, K.C.R.; Reshmi, R. Room temperature ammonia sensing of α-MoO3 nanorods grown on glass substrates. Thin Solid Film. 2021, 722, 138575–138584. [Google Scholar] [CrossRef]
  48. Shen, S.; Zhang, X.; Cheng, X.; Xu, Y.; Gao, S.; Zhao, H.; Zhou, X.; Huo, L. Oxygen-Vacancy-Enriched Porous α-MoO3 Nanosheets for Trimethylamine Sensing. ACS Appl. Nano Mater. 2019, 2, 8016–8026. [Google Scholar] [CrossRef]
  49. Zhang, J.; Song, P.; Li, Z.; Zhang, S.; Yang, Z.; Wang, Q. Enhanced trimethylamine sensing performance of single-crystal MoO3 nanobelts decorated with Au nanoparticles. J. Alloys Compd. 2016, 685, 1024–1033. [Google Scholar] [CrossRef]
  50. He, S.; Li, W.; Feng, L.; Yang, W. Rational interaction between the aimed gas and oxide surfaces enabling high-performance sensor: The case of acidic α-MoO3 nanorods for selective detection of triethylamine. J. Alloys Compd. 2019, 783, 574–582. [Google Scholar] [CrossRef]
  51. Ma, Y.T.; Ma, S.Y.; Tang, J.; Wu, Z.G.; Shi, J.; Zhao, Y.; Wang, Y.; Pei, S.T. Constructed heterostructured SnS@MoO3 hollow nanotubes and detected sensing properties towards TEA. Vacuum 2021, 184, 109939–109943. [Google Scholar] [CrossRef]
  52. Fu, X.; Yang, P.; Xiao, X.; Zhou, D.; Huang, R.; Zhang, X.; Cao, F.; Xiong, J.; Hu, Y.; Tu, Y.; et al. Ultra-fast and highly selective room-temperature formaldehyde gas sensing of Pt-decorated MoO3 nanobelts. J. Alloys Compd. 2019, 797, 666–675. [Google Scholar] [CrossRef]
  53. Yu, H.; Wang, C.; Meng, F.-Y.; Liang, J.-G.; Kashan, H.S.; Adhikari, K.K.; Wang, L.; Kim, E.-S.; Kim, N.-Y. Design and analysis of ultrafast and high-sensitivity microwave transduction humidity sensor based on belt-shaped MoO3 nanomaterial. Sens. Actuators B Chem. 2020, 304, 127138–127147. [Google Scholar] [CrossRef]
  54. Mohamed Azharudeen, A.; Karthiga, R.; Rajarajan, M.; Suganthi, A. Selective enhancement of non-enzymatic glucose sensor by used PVP modified on α-MoO3 nanomaterials. Microchem. J. 2020, 157, 105006–105014. [Google Scholar] [CrossRef]
  55. Zeng, Q.; Chen, J.; Gao, F.; Tu, X.; Qian, Y.; Yu, Y.; Lu, L.; Wang, W. Development of a new electrochemical sensing platform based on MoO3-polypyrrole nanowires/MWCNTs composite and its application to luteolin detection. Synth. Met. 2020, 271, 116620–116627. [Google Scholar] [CrossRef]
  56. Rahman, F.; Zavabeti, A.; Rahman, M.A.; Arash, A.; Mazumder, A.; Walia, S.; Sriram, S.; Bhaskaran, M.; Balendhran, S. Dual Selective Gas Sensing Characteristics of 2D alpha-MoO3−x via a Facile Transfer Process. ACS Appl. Mater. Interfaces 2019, 11, 40189–40195. [Google Scholar] [CrossRef]
  57. Guo, J.; Li, H.; Ma, Q.; Zhang, Q.; Wang, Y.; Wang, X. Surface oxygen vacancies actuated MoO3/CuMoO4 self-assembled microspheres for highly selective triethylamine detection. Sens. Actuators B Chem. 2022, 369, 132256–132266. [Google Scholar] [CrossRef]
  58. Zheng, F.; Dong, F.; Lv, Z.; Li, H.; Zhou, L.; Chen, Y.; Huo, T.; Luo, X. A novel g-C3N4/tourmaline composites equipped with plasmonic MoO3−x to boost photocatalytic activity. Colloids Interface Sci. Commun. 2021, 43, 100434–100443. [Google Scholar] [CrossRef]
  59. Sui, L.-L.; Zhang, X.-F.; Cheng, X.; Wang, P.; Xu, Y.-M.; Gao, S.; Zhao, H.; Huo, L.-H. Au-loaded hierarchical MoO3 hollow spheres with enhanced gas sensing performance for the detection of BTX (benzene, toluene and xylene) and the sensing mechanism. ACS Appl. Mater. Interfaces 2017, 9, 1661–1670. [Google Scholar] [CrossRef]
  60. Yang, W.; Fang, B.; Xiao, X.; Meng, H.; Liu, S. Hierarchical core-shell heterostructures of α-MoO3 nanorods@NiO nanosheets for significant detection of ethyl acetate vapor. Sens. Actuators B Chem. 2022, 358, 131457–131466. [Google Scholar] [CrossRef]
  61. Chang, J.; Zhang, H.; Cao, J.; Wang, Y. Ultrahigh sensitive and selective triethylamine sensor based on h-BN modified MoO3 nanowires. Adv Powder Technol. 2022, 33, 103432–103443. [Google Scholar] [CrossRef]
  62. Oh, S.H.; Park, S.M.; Kang, D.-W.; Kang, Y.C.; Cho, J.S. Fibrous network of highly integrated carbon nanotubes/MoO3 composite bundles anchored with MoO3 nanoplates for superior lithium ion battery anodes. J. Ind. Eng. Chem. 2020, 83, 438–448. [Google Scholar] [CrossRef]
  63. Lee, H.-J.; Shim, H.-W.; Kim, J.-C.; Kim, D.-W. Mo-MoO3-graphene nanocomposites as anode materials for lithium-ion batteries: Scalable, facile preparation and characterization. Electrochim. Acta 2017, 251, 81–90. [Google Scholar] [CrossRef]
  64. Deng, L.; Wei, T.; Liu, J.; Zhan, L.; Chen, W.; Cao, J. Recent developments of carbon-based anode materials for flexible lithium-ion batteries. Crystals 2022, 12, 1279. [Google Scholar] [CrossRef]
  65. Pavlovskii, A.A.; Pushnitsa, K.; Kosenko, A.; Novikov, P.; Popovich, A.A. Organic Anode Materials for Lithium-Ion Batteries: Recent Progress and Challenges. Materials 2022, 16, 177. [Google Scholar] [CrossRef] [PubMed]
  66. Meng, L.; Peng, J.; Zhang, Y.; Cui, Y.; An, L.; Chen, P.; Zhang, F. Lithium vanadium oxide/graphene composite as a promising anode for lithium-ion batteries. Nanomaterials 2022, 13, 43. [Google Scholar] [CrossRef] [PubMed]
  67. Ding, X.; Liang, D.; Zhao, H. Enhanced electrochemical Performance Promoted by Tin in silica Anode materials for Stable and High-capacity Lithium-ion Batteries. Materials 2021, 14, 1071. [Google Scholar] [CrossRef] [PubMed]
  68. Yue, Y.; Hong, L. Micro- and Nano-Structured Vanadium Pentoxide (V2O5) for Electrodes of Lithium-Ion Batteries. Adv. Energy Mater. 2017, 7, 1602545–1602577. [Google Scholar] [CrossRef]
  69. Cui, Y.; Pu, Y.; Hao, Y.; Zhuang, Q. Electrochemical behavior of α-MoO3 nanobelts as cathode material for Lithium ions batteries Russ. J. Electrochem. 2015, 51, 119–124. [Google Scholar]
  70. Wang, X.J.; Nesper, R.; Villevieille, C.; Novák, P. Ammonolyzed MoO3 nanobelts as novel cathode material of rechargeable Li-Ion batteries. Adv. Energy Mater. 2013, 3, 606–614. [Google Scholar] [CrossRef]
  71. Opra, D.P.; Gnedenkov, S.V.; Sokolov, A.A.; Podgorbunsky, A.B.; Laptash, N.M.; Sinebryukhov, S.L. Fluorine substituted molybdenum oxide as cathode material for Li-ion battery. Mater. Lett. 2015, 160, 175–178. [Google Scholar] [CrossRef]
  72. Hashem, A.M.; Groult, H.; Mauger, A.; Zaghib, K.; Julien, C.M. Electrochemical properties of nanofibers a-MoO3 as cathode materials for Li batteries. J. Power Sources 2012, 219, 126–132. [Google Scholar] [CrossRef]
  73. Vincent, R.C.; Cheetham, A.K.; Seshadri, R. Structure and lithium insertion in oxides of molybdenum. APL Mater. 2023, 11, 010902–010912. [Google Scholar] [CrossRef]
  74. Ramana, C.; Mauger, A.; Julien, C.M. Growth, characterization and performance of bulk and nanoengineered molybdenum oxides for electrochemical energy storage and conversion. Prog. Cryst. Growth Charact. Mater. 2021, 67, 100533–100594. [Google Scholar] [CrossRef]
  75. Du, Y.; He, J.; Hou, G.; Yuan, F. α-MoO3 sheets with high exposed plane reinforced by thermal plasma for stable Li-ion storage. Electrochim. Acta 2020, 334, 135593–135602. [Google Scholar] [CrossRef]
  76. Hou, X.; Ruan, M.; Zhou, L.; Wu, J.; Meng, B.; Huang, W.; Zhong, K.; Yang, K.; Fang, Z.; Xie, K. Superior lithium storage performance in MoO3 by synergistic effects: Oxygen vacancies and nanostructures. J. Energy Chem. 2023, 78, 91–101. [Google Scholar] [CrossRef]
  77. Cao, L.; He, J.; Li, J.; Yan, J.; Huang, J.; Qi, Y.; Feng, L. Surface tiny grain-dependent enhanced rate performance of MoO3 nanobelts with pseudocapacitance contribution for lithium-ion battery anode. J. Power Sources 2018, 392, 87–93. [Google Scholar] [CrossRef]
  78. Paste, R.; Abbas, S.A.; Singh, A.; Lin, H.-C.; Chu, C.W. Oxygen-Enriched α-MoO3–x nanobelts suppress lithium dendrite formation in stable lithium-metal batteries. J. Power Sources 2021, 507, 230306–230316. [Google Scholar] [CrossRef]
  79. Chen, X.; Huang, Y.; Zhang, K. a-MoO3 nanorods coated with SnS2 nano sheets core-shell composite as high-performance anode materials of lithium ion batteries. Electrochim. Acta 2016, 222, 956–964. [Google Scholar] [CrossRef]
  80. Byeon, A.; Hatter, C.; Park, J.H.; Ahn, C.W.; Gogotsi, Y.; Lee, J.W. Molybdenum oxide/carbon composites derived from the CO2 oxidation of Mo2CTx (MXene) for lithium ion battery anodes. Electrochim. Acta 2017, 258, 979–987. [Google Scholar] [CrossRef]
  81. Cheng, X.; Li, Y.; Sang, L.; Ma, J.; Shi, H.; Liu, X.; Lu, J.; Zhang, Y. Boosting the electrochemical performance of MoO3 anode for long-life lithium ion batteries: Dominated by an ultrathin TiO2 passivation layer. Electrochim. Acta 2018, 269, 241–249. [Google Scholar] [CrossRef]
  82. Teng, Y.; Zhao, H.; Zhang, Z.; Li, Y.; Liu, H. Self-assembly of flower-like MoO3-NiO microspheres with carbon coating as high-performance anode material for lithium-ion batteries. Mater. Lett. 2019, 246, 141–143. [Google Scholar] [CrossRef]
  83. Wang, H.; Xie, S.; Yao, T.; Wang, J.; She, Y.; Shi, J.-W.; Shan, G.; Zhang, Q.; Han, X.; Leung, M.K. Casting amorphorized SnO2/MoO3 hybrid into foam-like carbon nanoflakes towards high-performance pseudocapacitive lithium storage. J. Colloid Interface Sci. 2019, 547, 299–308. [Google Scholar] [CrossRef] [PubMed]
  84. Naresh, N.; Jena, P.; Satyanarayana, N. Compounds, Facile synthesis of MoO3/rGO nanocomposite as anode materials for high performance lithium-ion battery applications. J. Alloys Compd. 2019, 810, 151920–151928. [Google Scholar] [CrossRef]
  85. Zheng, Y.; Liu, Z.; Liu, B.; Wang, S.; Xiong, C. Fabrication of cactus-like CNT/SiO2/MoO3 ternary composites for superior lithium storage. Energy 2021, 217, 119386–119393. [Google Scholar] [CrossRef]
  86. Liu, J.; Li, G.; Wu, J. Fe2O3–TeO2–MoO3 semiconductor glass-ceramics as anode materials for high specific capacity lithium ion batteries. Mater. Chem. Phys. 2021, 258, 123894–123903. [Google Scholar] [CrossRef]
  87. Feng, Y.; Wu, K.; Ke, J.; Guo, Z.; Deng, X.; Bai, C.; Sun, Y.; Wang, Q.; Yang, B.; Dong, H.; et al. Synthesis of ternary SnO2–MoO3–C composite with nanosheet structure as high-capacity, high-rate and long-lifetime anode for lithium-ion batteries. Ceram. Int. 2021, 47, 9303–9309. [Google Scholar] [CrossRef]
  88. Almodóvar, P.; López, M.L.; Ramírez-Castellanos, J.; Nappini, S.; Magnano, E.; González-Calbet, J.M.; Díaz-Guerra, C. Synthesis, characterization and electrochemical assessment of hexagonal molybdenum trioxide (h-MoO3) micro-composites with graphite, graphene and graphene oxide for lithium ion batteries. Electrochim. Acta 2021, 365, 137355–137367. [Google Scholar] [CrossRef]
  89. Deng, X.; Zhu, M.; Ke, J.; Feng, Y.; Li, W.; Xiong, D.; He, M. Synthesis and electrochemical performances of ternary nanocomposite SnO2@ MoO3@ graphene as high-performance anode material for lithium-ion batteries. Chem. Phys. Lett. 2021, 770, 138408–138417. [Google Scholar] [CrossRef]
  90. Deng, X.; Zhu, M.; Ke, J.; Li, W.; Feng, Y.; Xiong, D.; He, M. SnO2-MoO3 nanoparticles anchored in carbon nanotubes as a large-capacity, high-rate, and long-lifetime anode for lithium-ion batteries. Ceram. Int. 2021, 47, 27022–27031. [Google Scholar] [CrossRef]
  91. Shahsank, M.; Naik, H.B.; Sumedha, H.; Nagaraju, G. Implementing an in-situ carbon formation of MoO3 nanoparticles for high performance lithium-ion battery. Ceram. Int. 2021, 47, 10261–10267. [Google Scholar] [CrossRef]
  92. Chen, Y.; Zong, W.; Chen, H.; Li, Z.; Pang, H.; Yuan, A.; Yang, H.; Shen, X. Cyanide-metal framework derived porous MoO3-Fe2O3 hybrid micro-octahedrons as superior anode for lithium-ion batteries. Chem. Eng. J. 2021, 426, 130347–130354. [Google Scholar] [CrossRef]
  93. Zhang, P.; Cai, B.; Feng, Y.; Pan, H.; Yao, J. Constructing MoO3@ MoO2 heterojunction on g-C3N4 nanosheets with advanced Li-ion storage ability. J. Alloys Compd. 2021, 875, 160077–160084. [Google Scholar] [CrossRef]
  94. Wu, Y.; Cheng, X.; Zhang, X.; Xu, Y.; Gao, S.; Zhao, H.; Huo, L. High efficient and selective removal of Pb2+ through formation of lead molybdate on α-MoO3 porous nanosheets array. J. Colloid Interface Sci. 2017, 491, 80–88. [Google Scholar] [CrossRef] [PubMed]
  95. Sulejmanović, J.; Kovač, N.; Memić, M.; Šabanović, E.; Begić, S.; Sher, F.; Engineering, E. Selective removal of lead ions from aqueous solutions using SiO2–MoO3: Isotherm, kinetics and thermodynamic studies. Case Stud. Chem. Environ. Eng. 2021, 3, 100083–100087. [Google Scholar] [CrossRef]
  96. Chen, J.; Liao, Y.; Wan, X.; Tie, S.; Zhang, B.; Lan, S.; Gao, X. A high performance MoO3@ MoS2 porous nanorods for adsorption and photodegradation of dye. J. Solid State Chem. 2020, 291, 121652–121662. [Google Scholar] [CrossRef]
  97. Zhou, W.; Deng, J.; Qin, Z.; Huang, R.; Wang, Y.; Tong, S. Construction of MoS2 nanoarrays and MoO3 nanobelts: Two efficient adsorbents for removal of Pb (II), Au (III) and Methylene Blue. J. Environ. Sci. 2022, 111, 38–50. [Google Scholar] [CrossRef]
  98. Kumar, N.; Kumar, R. Physics, Efficient adsorption of methylene blue on hybrid structural phase of MoO3 nanostructures. Mater. Chem. Phys. 2022, 275, 125211–125225. [Google Scholar] [CrossRef]
  99. Ren, P.; Ren, X.; Xu, J.; Li, H.; Zheng, Y.; Hong, Y.; Lin, Y.; Zhou, Y.; Chen, Y.; Zhang, W. Excellent adsorption property and mechanism of oxygen vacancies-assisted hexagonal MoO3 nanosheets for methylene blue and rhodamine b dyes. Appl. Surf. Sci. 2022, 597, 153699–153709. [Google Scholar] [CrossRef]
  100. Ma, Y.; Jia, Y.; Jiao, Z.; Wang, L.; Yang, M.; Bi, Y.; Qi, Y. Facile synthesize α-MoO3 nanobelts with high adsorption property. Mater. Lett. 2015, 157, 53–56. [Google Scholar] [CrossRef]
  101. Kedves, E.-Z.; Bárdos, E.; Gyulavári, T.; Pap, Z.; Hernadi, K.; Baia, L. Dependence of cationic dyes’ adsorption upon α-MoO3 structural properties. Appl. Surf. Sci. 2022, 573, 151584–151596. [Google Scholar] [CrossRef]
  102. Zhao, Y.; Sun, Q.; Luo, J.; Chen, H.; Cai, W.; Su, X. Hydrothermal fabrication of TiO2-MoO3 nanocomposites with superior performance for water treatment. Nano-Struct. Nano-Objects 2018, 13, 93–99. [Google Scholar] [CrossRef]
  103. Muthukumar, P.; Sowmiya, E.; Arunkumar, G.; Pannipara, M.; Al-Sehemi, A.G.; Anthony, S.P. Highly enhanced dye adsorption of MoO3 nanoplates fabricated by hydrothermal-calcination approach in presence of chitosan and thiourea. Chemosphere 2022, 291, 132926–132936. [Google Scholar] [CrossRef]
  104. Heiba, H.F.; Taha, A.A.; Mostafa, A.R.; Mohamed, L.A.; Fahmy, M.A. Preparation and characterization of novel mesoporous chitin blended MoO3-montmorillonite nanocomposite for Cu (II) and Pb (II) immobilization. Int. J. Biol. Macromol. 2020, 152, 554–566. [Google Scholar] [CrossRef]
  105. Allam, E.A.; El-Sharkawy, R.M.; Gizawy, M.; Mahmoud, M.E. Physics, Assembly of CeO2–MoO3–SiO2 (CH2) 3-(Alginate)2 As A novel nanocomposite for removal of MnII/CrVI and 56Mn/51Cr radionuclides from water. Mater. Chem. Phys. 2021, 262, 124278–124293. [Google Scholar] [CrossRef]
  106. Xue, L.; Zhang, D.; Xu, Y.; Liu, X.; Materials, M. Adsorption of thiophene compounds on MoO3/γ-Al2O3 catalysts with different mesopore sizes. Microporous Mesoporous Mater. 2017, 238, 46–55. [Google Scholar] [CrossRef]
  107. Wang, Q.; Tian, H.; Zhang, Z.; Dang, T.; Zhang, W.; Wang, J.; Lu, Y.; Liu, S. Keggin-type polycationic AlO4Al12 (OH)24 (H2O) 127+ intercalated MoO3 composites for methyl orange adsorption. Chin. Chem. Lett. 2022, 33, 2617–2620. [Google Scholar] [CrossRef]
  108. Huang, J.; Ye, M.; Qu, Y.; Chu, L.; Chen, R.; He, Q.; Xu, D. Pb (II) removal from aqueous media by EDTA-modified mesoporous silica SBA-15. J. Colloid Interface Sci. 2012, 385, 137–146. [Google Scholar] [CrossRef] [PubMed]
  109. Li, Y.; Huang, L.; Xu, J.; Xu, H.; Xu, Y.; Xia, J.; Li, H. Visible-light-induced blue MoO3–C3N4 composite with enhanced photocatalytic activity. Mater. Res. Bull. 2015, 70, 500–505. [Google Scholar] [CrossRef]
  110. Saravanakumar, K.; Balakumar, V.; Jones, B.F.; Muthuraj, V. Physics, A structural unique 1D-MoO3@ 3D-WO3 nanohybrid for stable and reusable photocatalytic conversion of hexavalent chromium in aqueous medium. Mater. Chem. Phys. 2021, 267, 124688–124699. [Google Scholar] [CrossRef]
  111. Xiao, J.; Lv, J.; Lu, Q. Building Fe2O3/MoO3 nanorod heterojunction enables better tetracycline photocatalysis. Mater. Lett. 2022, 311, 131580–131584. [Google Scholar] [CrossRef]
  112. Li, Y.; Chen, C.; Chen, X.; Zang, J. MoO3/g-C3N4 heterostructure for degradation of organic pollutants under visible light irradiation: High efficiency, general degradation and Z-scheme degradation mechanism. Ceram. Int. 2021, 47, 33697–33708. [Google Scholar] [CrossRef]
  113. Agarwal, P.; Paramasivam, I.; Shrestha, N.K.; Schmuki, P. MoO3 in self-organized TiO2 nanotubes for enhanced photocatalytic activity. Chem. Asian J. 2010, 5, 66–69. [Google Scholar] [CrossRef] [PubMed]
  114. Li, N.; Li, Y.; Zhou, Y.; Li, W.; Ji, S.; Yao, H.; Cao, X.; Jin, P. Interfacial-charge-transfer-induced photochromism of MoO3@ TiO2 crystalline-core amorphous-shell nanorods. Sol. Energy Mater. Sol. Cells 2017, 160, 116–125. [Google Scholar] [CrossRef]
  115. Arfaoui, A.; Mhamdi, A.; Besrour, N.; Touihri, S.; Ouzari, H.; Alrowaili, Z.; Amlouk, M. Investigations into the physical properties of SnO2/MoO3 and SnO2/WO3 bi-layered structures along with photocatalytic and antibacterial applications. Thin Solid Film. 2018, 648, 12–20. [Google Scholar] [CrossRef]
  116. Patil, S.B.; Kishore, B.; Manjunath, K.; Reddy, V.; Nagaraju, G. One step hydrothermal synthesis of novel Cu2S-MoO3 nanocomposite for lithium ion battery and photocatalytic applications. Int. J. Hydrogen Energy 2018, 43, 4003–4014. [Google Scholar] [CrossRef]
  117. Zhu, J.-M.; Hosseini, M.; Fakhri, A.; Rad, S.S.; Hadadi, T.; Nobakht, N.; Biology, P.B. Highly efficient of molybdenum trioxide-cadmium titanate nanocomposites for ultraviolet light photocatalytic and antimicrobial application: Influence of reactive oxygen species. J. Photochem. Photobiol. B Biol. 2019, 191, 75–82. [Google Scholar] [CrossRef] [PubMed]
  118. Sharma, S.; Basu, S. Visible-light-driven efficient photocatalytic abatement of recalcitrant pollutants by centimeter-length MoO3/SiO2 monoliths with long service life. Appl. Mater. Today 2021, 23, 101033–101045. [Google Scholar] [CrossRef]
  119. Alenazi, D.A.K.; Aslam, M.; Chandrasekaran, S.; Soomro, M.T.; Ali, S.; Danish, E.Y.; Ismail, I.; Hameed, A. Facile fabrication of MoO3/g-C3N4 pn junction for boosted photocatalytic elimination of 2,4-D under natural sunlight exposure. J. Environ. Chem. Eng. 2021, 9, 106322. [Google Scholar] [CrossRef]
  120. Li, J.; Liu, H.; Liu, Z.; Yang, D.; Zhang, M.; Gao, L.; Zhou, Y.; Lu, C. Facile synthesis of Z-scheme NiO/α-MoO3 pn heterojunction for improved photocatalytic activity towards degradation of methylene blue. Arab. J. Chem. 2022, 15, 103513–103525. [Google Scholar] [CrossRef]
  121. Srinithi, S.; Balakumar, V.; Chen, S.-M. In-situ fabrication of polypyrrole composite with MoO3: An effective interfacial charge transfers and electrode materials for degradation and determination of acetaminophen. Chemosphere 2022, 291, 132977–132987. [Google Scholar] [CrossRef]
  122. Paul, M.; Dhanasekar, M.; Bhat, S.V. Silver doped h-MoO3 nanorods for sonophotocatalytic degradation of organic pollutants in ambient sunlight. Appl. Surf. Sci. 2017, 418, 113–118. [Google Scholar] [CrossRef]
  123. Liu, Q.; Wu, Y.; Zhang, J.; Chen, K.; Huang, C.; Chen, H.; Qiu, X. Plasmonic MoO3−x nanosheets with tunable oxygen vacancies as efficient visible light responsive photocatalyst. Appl. Surf. Sci. 2019, 490, 395–402. [Google Scholar] [CrossRef]
  124. Li, J.; Xu, X.; Huang, B.; Lou, Z.; Li, B. Light-induced in situ formation of a nonmetallic plasmonic MoS2/MoO3–x heterostructure with efficient charge transfer for CO2 reduction and SERS detection. ACS Appl. Mater. Interfaces 2021, 13, 10047–10053. [Google Scholar] [CrossRef] [PubMed]
  125. Gao, Q.; Wang, Z.; Rao, Y.; Zhao, Y.; Cao, J.; Ho, K.-F.; Zhai, Y.; Xiong, M.; Li, J.; Huang, Y. Oxygen vacancy mediated α-MoO3 bactericidal nanocatalyst in the dark: Surface structure dependent superoxide generation and antibacterial mechanisms. J. Hazard. Mater. 2023, 443, 130275–130285. [Google Scholar] [CrossRef] [PubMed]
  126. Yang, Y.-Y.; Niu, C.-G.; Huang, D.-W.; Guo, H.; Feng, H.-P.; Li, L.; Liu, H.-Y.; Fan, Q.-Q.; Qin, M.-Z. Appropriate oxygen vacancies and Mo-N bond synergistically modulate charge transfer dynamics of MoO3−x/S-CN for superior photocatalytic disinfection: Unveiling synergistic effects and disinfection mechanism. J. Hazard. Mater. 2023, 445, 130481–130496. [Google Scholar] [CrossRef]
  127. Wang, K.-L.; Zhao, H.-M.; Li, Y.; Sun, T.; Xue, B. Highly dispersed molybdenum-embedded polymeric carbon nitride with enhanced photocatalytic activity for degradation of dyes and antibiotics. Appl. Surf. Sci. 2020, 528, 146931–146938. [Google Scholar] [CrossRef]
  128. Sharma, P.K.; Raghubanshi, A.; Shah, K. Examining dye degradation and antibacterial properties of organically induced α-MoO3 nanoparticles, their uptake and phytotoxicity in rice seedlings. Environ. Nanotechnol. Monit. Manag. 2020, 14, 100315–100325. [Google Scholar]
  129. Sales, H.B.; Menezes, R.R.; Neves, G.A.; Souza, J.J.d.; Ferreira, J.M.; Chantelle, L.; Menezes de Oliveira, A.L.; Lira, H.L. Development of sustainable heterogeneous catalysts for the photocatalytic treatment of effluents. Sustainability 2020, 12, 7393. [Google Scholar] [CrossRef]
  130. Hussain, M.K.; Khalid, N.R. Surfactant-assisted synthesis of MoO3 nanorods and its application in photocatalytic degradation of different dyes in aqueous environment. J. Mol. Liq. 2022, 346, 117871–117882. [Google Scholar] [CrossRef]
  131. Dewangan, K.; Singh, D.; Satpute, N.; Singh, R.; Jaiswal, A.; Shrivas, K.; Bahadur, I. Hydrothermally grown α-MoO3 microfibers for photocatalytic degradation of methylene blue dye. J. Mol. Liq. 2022, 349, 118202–118212. [Google Scholar] [CrossRef]
  132. Kader, S.; Al-Mamun, M.R.; Suhan, M.B.K.; Shuchi, S.B.; Islam, M.S. Innovation, Enhanced photodegradation of methyl orange dye under UV irradiation using MoO3 and Ag doped TiO2 photocatalysts. Environ. Technol. Innov. 2022, 27, 102476–102491. [Google Scholar] [CrossRef]
  133. Zhang, Y.; Park, S.-J. Facile construction of MoO3@ ZIF-8 core-shell nanorods for efficient photoreduction of aqueous Cr (VI). Appl. Catal. B 2019, 240, 92–101. [Google Scholar] [CrossRef]
  134. Al-Namshah, K.; Mohamed, R. Decoration of MoO3 nanoparticles by MWCNTs driven visible light for the reduction of Cr (VI). Ceram. Int. 2020, 46, 6914–6919. [Google Scholar] [CrossRef]
Figure 1. Published scientific papers directly related to MoO3 and MoO3-based systems in the last 20 years (Web of Science search with MoO3 keyword, 2023 ongoing, Access on 9 October 2023).
Figure 1. Published scientific papers directly related to MoO3 and MoO3-based systems in the last 20 years (Web of Science search with MoO3 keyword, 2023 ongoing, Access on 9 October 2023).
Materials 16 07657 g001
Figure 2. MoO3 crystalline structure illustrations.
Figure 2. MoO3 crystalline structure illustrations.
Materials 16 07657 g002
Figure 3. Right panel: Crystal structure of MoO3 with H+ intercalation and hydrogen molybdenum bronze. Left panel: Electronic band structure manipulation by H+ intercalation. Figure reproduced with permission of [24] Copyright 2017, Nature Publishing Group.
Figure 3. Right panel: Crystal structure of MoO3 with H+ intercalation and hydrogen molybdenum bronze. Left panel: Electronic band structure manipulation by H+ intercalation. Figure reproduced with permission of [24] Copyright 2017, Nature Publishing Group.
Materials 16 07657 g003
Figure 4. Scheme of the reaction mechanism of MoO3 materials exposed to hydrogen gas. Based on Ref. [36].
Figure 4. Scheme of the reaction mechanism of MoO3 materials exposed to hydrogen gas. Based on Ref. [36].
Materials 16 07657 g004
Figure 5. Published scientific papers and citations directly related to MoO3 and MoO3-based systems applied to gas sensors in the last 20 years (Web of Science search with MoO3 and gas sensors keywords, 2023 ongoing, Access on 9 October 2023).
Figure 5. Published scientific papers and citations directly related to MoO3 and MoO3-based systems applied to gas sensors in the last 20 years (Web of Science search with MoO3 and gas sensors keywords, 2023 ongoing, Access on 9 October 2023).
Materials 16 07657 g005
Figure 6. Responses of MoO3 nanorods to NO2, CO, and CH4 gases at 40 ppm and different temperatures. Figure reproduced with permission of [41] Copyright 2012, Elsevier.
Figure 6. Responses of MoO3 nanorods to NO2, CO, and CH4 gases at 40 ppm and different temperatures. Figure reproduced with permission of [41] Copyright 2012, Elsevier.
Materials 16 07657 g006
Figure 7. Schematic diagram illustrating the 1-butylamine sensing mechanism on the MoO3 surface showing the dehydrogenation pathway of (a) 1-butylamine and (b) 1-butylimine. Figure reproduced with permission of [37] Copyright 2021, Elsevier.
Figure 7. Schematic diagram illustrating the 1-butylamine sensing mechanism on the MoO3 surface showing the dehydrogenation pathway of (a) 1-butylamine and (b) 1-butylimine. Figure reproduced with permission of [37] Copyright 2021, Elsevier.
Materials 16 07657 g007
Figure 8. A schematic diagram illustrates an external load applied to both electrodes. This is the discharging process of a lithium-ion battery. Figure reproduced with permission of [68] Copyright 2017, Wiley.
Figure 8. A schematic diagram illustrates an external load applied to both electrodes. This is the discharging process of a lithium-ion battery. Figure reproduced with permission of [68] Copyright 2017, Wiley.
Materials 16 07657 g008
Figure 9. Published scientific papers and citations directly related to MoO3 and MoO3-based systems applied to LIBs in the last 20 years (Web of Science search with MoO3 and lithium-ion batteries keywords, 2023 ongoing, Access on 9 October 2023).
Figure 9. Published scientific papers and citations directly related to MoO3 and MoO3-based systems applied to LIBs in the last 20 years (Web of Science search with MoO3 and lithium-ion batteries keywords, 2023 ongoing, Access on 9 October 2023).
Materials 16 07657 g009
Figure 10. Published scientific papers and citations directly related to MoO3 and MoO3-based systems applied to adsorption in the last 20 years (Web of Science search with MoO3 and adsorption keywords, 2023 ongoing, Access on 9 October 2023).
Figure 10. Published scientific papers and citations directly related to MoO3 and MoO3-based systems applied to adsorption in the last 20 years (Web of Science search with MoO3 and adsorption keywords, 2023 ongoing, Access on 9 October 2023).
Materials 16 07657 g010
Figure 11. Removal efficiency of metal ions Cu2+, Pb2+, Zn2+, Cr3+, and Cd2+ removed by the α-MoO3 nanosheet array system. Figure reproduced with permission of [94] Copyright 2017, Elsevier.
Figure 11. Removal efficiency of metal ions Cu2+, Pb2+, Zn2+, Cr3+, and Cd2+ removed by the α-MoO3 nanosheet array system. Figure reproduced with permission of [94] Copyright 2017, Elsevier.
Materials 16 07657 g011
Figure 12. Published scientific papers and citations directly related to MoO3 and MoO3-based systems applied to photocatalysis in the last 20 years (Web of Science search with MoO3 and photocatalysis keywords, 2023 ongoing, Access on 9 October 2023).
Figure 12. Published scientific papers and citations directly related to MoO3 and MoO3-based systems applied to photocatalysis in the last 20 years (Web of Science search with MoO3 and photocatalysis keywords, 2023 ongoing, Access on 9 October 2023).
Materials 16 07657 g012
Figure 13. Schematic representation of visible light photocatalytic degradation of MB dye by MoO3.
Figure 13. Schematic representation of visible light photocatalytic degradation of MB dye by MoO3.
Materials 16 07657 g013
Table 3. MoO3 and MoO3-based systems used in adsorption applications.
Table 3. MoO3 and MoO3-based systems used in adsorption applications.
AdsorbentMorphologySynthesis Method Adsorbent NatureTargetqm (mg·g−1)Reference
α-MoO3 NanobeltsSolvothermalCationic dyeMB1408[97]
α-MoO3NanobeltsHydrothermalCationic dyeMBUD a[100]
α-MoO3Platelet-like crystalsCalcination Cationic dyeRhB26.56[101]
α-MoO3Platelet-like crystalsCalcination Cationic dyeMB152.7[101]
α-MoO3Platelet-like crystalsCalcination Cationic dyeCV199.4[101]
α-MoO3Platelet-like crystalsCalcination Cationic dyeMG53.05[101]
α-MoO3Micro/nanoplatesHydrothermalCationic dyeRhBUD a[103]
α-MoO3Micro/nanoplatesHydrothermalCationic dyeMBUD a[103]
h-MoO3NanosheetsHydrothermal Cationic dyeRhB1242[99]
h-MoO3NanosheetsHydrothermal Cationic dyeMB1433[99]
α-MoO3/h-MoO3NanoparticlesInternal combustion Cationic dyeMB141.2[98]
α-MoO3-MoO2NanoparticlesHydrothermalCationic dyeMB1250[22]
α-MoO3-MoS2Porous core-shell nanorodsHydrothermalCationic dyeRhB326.83[96]
α-MoO3-TiO2NanoparticlesHydrothermalCationic dyeRhB169[102]
α-MoO3-TiO2NanoparticlesHydrothermalCationic dyeMB180[102]
α-MoO3Micro/nanoplatesHydrothermalAnionic dyeEosin yellowUD a[103]
Al13-MoO3Flake-likePolycationic EncapsulationAnionic dyeMO357.2[107]
α-MoO3 Porous nanosheet arrayHydrothermalMetal ionPb2+1450[94]
α-MoO3 NanobeltsSolvothermalMetal ionPb2+684.93[97]
SiO2-α-MoO3UD aThermal decompositionMetal ionPb2+222.2[95]
α-MoO3-TiO2NanoparticlesHydrothermalMetal ionCr3+59[102]
Chitin-MoO3-MontmorilloniteNanorodsBlendingMetal ionCu2+19.03[104]
Chitin-MoO3-MontmorilloniteNanorodsBlendingMetal ionPb2+15.92[104]
CeO2–MoO3–SiO2(CH2)3-(Alginate)2Nanoparticle Combustion/microwave irradiation Metal ionMn2+122.06[105]
CeO2–MoO3–SiO2(CH2)3-(Alginate)2Nanoparticle Combustion/microwave irradiation Metal ionCr6+151.96[105]
MoO3/γ-Al2O3Nanoparticles/NanorodsCation-anion double hydrolysis/ImpregnationAromatic sulfur compoundDibenzothiopheneUD a[106]
MoO3/γ-Al2O3Nanoparticles/NanorodsCation-anion double hydrolysis/ImpregnationAromatic sulfur compoundThiopheneUD a[106]
a Unavailable data.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

da Silva Júnior, M.G.; Arzuza, L.C.C.; Sales, H.B.; Farias, R.M.d.C.; Neves, G.d.A.; Lira, H.d.L.; Menezes, R.R. A Brief Review of MoO3 and MoO3-Based Materials and Recent Technological Applications in Gas Sensors, Lithium-Ion Batteries, Adsorption, and Photocatalysis. Materials 2023, 16, 7657. https://doi.org/10.3390/ma16247657

AMA Style

da Silva Júnior MG, Arzuza LCC, Sales HB, Farias RMdC, Neves GdA, Lira HdL, Menezes RR. A Brief Review of MoO3 and MoO3-Based Materials and Recent Technological Applications in Gas Sensors, Lithium-Ion Batteries, Adsorption, and Photocatalysis. Materials. 2023; 16(24):7657. https://doi.org/10.3390/ma16247657

Chicago/Turabian Style

da Silva Júnior, Mário Gomes, Luis Carlos Costa Arzuza, Herbet Bezerra Sales, Rosiane Maria da Costa Farias, Gelmires de Araújo Neves, Hélio de Lucena Lira, and Romualdo Rodrigues Menezes. 2023. "A Brief Review of MoO3 and MoO3-Based Materials and Recent Technological Applications in Gas Sensors, Lithium-Ion Batteries, Adsorption, and Photocatalysis" Materials 16, no. 24: 7657. https://doi.org/10.3390/ma16247657

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

Article Metrics

Back to TopTop