Next Article in Journal
Layer Morphology and Ink Compatibility of Silver Nanoparticle Inkjet Inks for Near-Infrared Sintering
Next Article in Special Issue
Bifunctional Tm3+,Yb3+:GdVO4@SiO2 Core-Shell Nanoparticles in HeLa Cells: Upconversion Luminescence Nanothermometry in the First Biological Window and Biolabelling in the Visible
Previous Article in Journal
Artificial Intelligence Based Methods for Asphaltenes Adsorption by Nanocomposites: Application of Group Method of Data Handling, Least Squares Support Vector Machine, and Artificial Neural Networks
Previous Article in Special Issue
Coulomb Blockade Effect in Well-Arranged 2D Arrays of Palladium Nano-Islands for Hydrogen Detection at Room Temperature: A Modeling Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Hydrothermal Synthesis and Gas Sensing of Monoclinic MoO3 Nanosheets

by
Teodóra Nagyné-Kovács
1,
Levente Studnicka
1,
István Endre Lukács
2,
Krisztina László
3,
Pawel Pasierb
4,
Imre Miklós Szilágyi
1,* and
György Pokol
1,5
1
Department of Inorganic and Analytical Chemistry, Budapest University of Technology and Economics, Műegyetem rakpart 3, H-1111 Budapest, Hungary
2
Research Institute for Technical Physics and Materials Science, Eötvös Loránd Research Network, Konkoly Thege M. út 29-33, H-1121 Budapest, Hungary
3
Department of Physical Chemistry and Materials Science, Budapest University of Technology and Economics, Műegyetem rakpart 3, H-1111 Budapest, Hungary
4
Faculty of Materials Science and Ceramics, AGH University of Science and Technology, Al. Mickiewicza 30, 30-059 Kraków, Poland
5
Research Centre for Natural Sciences, Eötvös Loránd Research Network, Magyar tudósok körútja 2, H-1117 Budapest, Hungary
*
Author to whom correspondence should be addressed.
Nanomaterials 2020, 10(5), 891; https://doi.org/10.3390/nano10050891
Submission received: 31 March 2020 / Revised: 24 April 2020 / Accepted: 27 April 2020 / Published: 7 May 2020
(This article belongs to the Special Issue Nanosensors)

Abstract

:
Effects of different reaction parameters in the hydrothermal synthesis of molybdenum oxides (MoO3) were investigated and monoclinic (β-) MoO3 was prepared hydrothermally for the first time. Various temperatures (90/210 °C, and as a novelty 240 °C) and durations (3/6 h) were used. At 240 °C, cetyltrimethylammonium bromide (CTAB) and CrCl3 additives were also tested. Both the reaction temperatures and durations played a significant role in the formation of the products. At 90 °C, h-MoO3 was obtained, while at 240 °C the orthorhombic (α-) MoO3 formed with hexagonal rod-like and nanofibrous morphology, respectively. The phase transformation between these two phases was observed at 210 °C. At this temperature, the 3 h reaction time resulted in the mixture of h- and α-MoO3, but 6 h led to pure α-MoO3. With CTAB the product was bare o-MoO3, however, when CrCl3 was applied, pure metastable m-MoO3 formed with the well-crystallized nanosheet morphology. The gas sensing of the MoO3 polymorphs was tested to H2, which was the first such gas sensing study in the case of m-WO3. Monoclinic MoO3 was found to be more sensitive in H2 sensing than o-MoO3. This initial gas sensing study indicates that m-MoO3 has promising gas sensing properties and this MoO3 polymorph is promising to be studied in detail in the future.

Graphical Abstract

1. Introduction

Molybdenum oxides (MoO3) are considerable materials among molybdenum (Mo) compounds due to their excellent physical and chemical properties. MoO3 is one of the most important starting materials for other Mo compounds, such as for sodium molybdate, ammonium di- or heptamolybdate, or the Mo metal itself. MoO3 is also used as a catalyst in the industrial production of acrylonitrile from propylene and ammonia. However, the constant interest in nanomaterials has made also nano MoO3 attractive, which has been proven to be a promising candidate in many fields, e.g., (photo)catalysis, gas sensing, electrochemical cells, or even in forensic science and photothermal therapy [1,2,3,4,5,6,7,8,9,10,11,12,13].
MoO3 has four allotropes, each composed of differently linked MoO6 octahedra, the position of which determines phase stability. The most commonly studied MoO3 phases are the thermodynamically stable orthorhombic (o-) or α-MoO3 and metastable monoclinic (m-) or β-MoO3. Both phases have excellent physical and chemical properties such as refractive index, band gap, and mechanical hardness. Further crystalline phases are the metastable high-pressure ε-MoO3 and the relatively more stable hexagonal (h-) MoO3 [14,15,16].
In the case of o-MoO3, octahedra are distorted and form bilayers, which build up the well-known stratified structure. These planar bilayers are composed of chains of MoO6 octahedra sharing corners and edges in the specified direction and are held together vertically by weak Van der Waals forces. In contrast, the crystal lattice of the metastable m-MoO3 is similar to the cubic ReO3, where not layered, but 3D crystal structure forms due to the corner-sharing of distorted octahedra in the three directions. The h-MoO3 phase is constructed by such chains in which octahedra are linked by their corners. These chains connect in such a way to form a hexagonal structure with the typical 1D hexagonal and trigonal channels. The hexagonal channels may contain small ions (e.g., alkali metal or ammonium ions) or water molecules, similar to hexagonal WO3 [14,15,16,17].
There are several reports about the hydrothermal preparation of MoO3, but mostly only o- and h-MoO3 are synthetized. Although the synthesis of the m-MoO3 phase has been carried out by many methods (spray pyrolysis, pulsed laser thin-layer deposition, molecular beam epitaxy, o-MoO3 conversion, solution-phase reaction, etc.), it has not been prepared hydrothermally yet in pure form, only as a part of a mixture of MoO3 polymorphs [18]. On the other hand, it would be useful because the hydrothermal method is simple, does not use high temperatures (usually it operates only between 100 and 250 °C), results in highly crystalline, monodisperse products, and metastable phases can be also synthetized by it. Based on the literature, the highest temperatures used in experiments to prepare MoO3 are typically 200–210 °C. However, it would be worth trying to use higher reaction temperatures as well, considering how important a small change in the reaction parameters can be during the hydrothermal synthesis in general, regarding the crystalline phase and morphology of the products. Consequently, neither the different reaction times nor the role of CTAB (cetyltrimethylammonium bromide) and CrCl3 additives have been studied yet at temperatures above 200–210 °C in the hydrothermal synthesis of MoO3. Furthermore, so far the gas sensing properties mostly of o- and h-MoO3 were tested [10,19,20,21,22,23,24,25,26,27,28,29,30]. To the best of our knowledge, there is only one paper reporting the gas sensing of m-MoO3, where very brief gas sensing data are given using H2O and CO2 as test gases [31].
In the present investigation, we studied different reaction parameters in the hydrothermal synthesis of MoO3, focusing on the changes in the obtained crystalline phases and morphologies. We investigated the effect of reaction temperature at 90, 210, and 240 °C (using as high as 240 °C for the first time), and duration applying 3 and 6 h reaction times. Finally, we tested CTAB and CrCl3 additives for the first time at 240 °C. We examined the crystalline phases and morphologies by X-ray diffraction (XRD) and scanning electron microscopy (SEM), respectively. The as-prepared different MoO3 allotropes were further studied by energy dispersive X-ray spectroscopy (EDX) and specific surface area measurements, moreover by Fourier transformed infrared (FT-IR), Raman, and ultraviolet-visible (UV–Vis) spectroscopies. Their band gap energies were also determined and their gas sensing properties were analyzed towards H2.

2. Materials and Methods

2.1. Materials and Equipment

Ammonium heptamolybdate tetrahydrate (≥99.0%), hexadecyltrimethylammonium bromide (≥96.0%), and chromium(III) chloride hexahydrate (≥98.0%) were purchased from Sigma-Aldrich (Budapest, Hungary) and used as received. Concentrated HNO3 (65%) was ordered from Molar Chemicals (Halásztelek, Hungary), while ethanol (96 V/V%) from WVR (Budapest, Hungary). Ion exchanged water was used throughout the experiments.
For the hydrothermal reactions a 45 mL stainless steel, Teflon-lined autoclave (4744, General Purpose Vessel, Parr Instruments, Moline, IL, USA), and a Nabertherm (Lilienthal, Germany) muffle-furnace (L9/11/B410) were used.

2.2. Hydrothermal Reactions

For every synthesis, 0.5 g of ammonium heptamolibdate (AHM) was dissolved in 25 mL of ion exchanged water, and then 5 mL cc. HNO3 was added. To study the effect of various additives, 0.1 g of CTAB or CrCl3 was also dissolved in the selected reactions. After stirring the solution for a couple of minutes, it was transferred into the autoclave, which was put into the electric furnace at particular temperatures and for different durations. After the hydrothermal treatment, the precipitates were filtered and washed several times with water and ethanol. Finally, the samples were dried at 60 °C for 2 h.
The hydrothermal reactions and their conditions are listed in Table 1.

2.3. Characterization

X-ray powder diffraction (XRD) measurements were carried out by a X’Pert Pro MPD diffractometer (PANalytical, Almelo, Netherlands) with Cu Kα radiation (λ = 0.15418 nm). For studying the morphology of the samples, a LEO 1540 XB electron microscope (Zeiss, Oberkochen, Germany) was used, while energy-dispersive X-ray spectroscopy (EDX) analyses were performed using a JEOL JSM 5500-LV instrument (JEOL, Tokyo, Japan). FT-IR spectra were recorded by a Perkin Elmer 2000 FT-IR spectrometer (Perkin Elmer, Boston, MA, USA) between 450 and 4000 cm−1 using KBr pellets (1 mg sample/300 mg KBr). Raman spectra were taken by a Jobin Yvon LabRam spectrometer (Horiba, Kyoto, Japan) equipped with an Olympus BX41 optical microscope (Olympus, Tokyo, Japan) using a frequency-doubled Nd-YAG laser (532 nm), while for diffuse reflectance UV–Vis measurements, a Jasco V-570 UV/ VIS/ NIR spectrometer (Jasco, Tokyo, Japan) was used. The band gaps of the products were calculated based on the UV–Vis spectra using αhυ = A (hυ − Eg)n equation, where α, , A, and Eg are equal to the molar absorption coefficient, photon energy, general constant, and band gap energy, respectively. The n depends on the type of the electron transition in the compound and is ½ (direct allowed) for MoO3. Using the Tauc-plot, which is plotting (αhυ)2 against , and then drawing a tangent line onto the linear range and extrapolating, the band gap energy can be determined. The Kubelka-Munk function was used to approximate the value of A. The N2 adsorption was measured at −196 °C with a Nova2000e (Quantachrome) computer-controlled apparatus and the apparent surface area (SBET) of the samples were calculated from the Brunauer–Emmett–Teller (BET) model [32].

2.4. Gas Sensing Tests

For the experiments, Al2O3 ceramic sensor plates with interdigital platinum electrodes were used. First, with the usage of gold paste gold strands were attached onto the electrodes. Then, the sensor plate was affixed to the end of a vertical ceramic tube, and it was connected to a circuit, to which an electrical resistance meter (HP 34410A) was connected. The samples were suspended in one drop of ethanol in an Eppendorf tube, and a small amount of the suspension was transferred to the electrode. For heating the sensor chips during the gas sensing measurements, an online, temperature, and atmosphere controlled furnace was utilized, which hosted the interdigital electrodes mounted on the vertical ceramic tube (see details in [33]). In the experiments, H2 gas with different concentrations (25,000/830/2500/5000/7500 and 10,000 ppm) was used and the carrier gas was N2. After the temperature was set, the gas mixture was introduced to the sample and the change in the electrical resistance of the sensor material was detected. The electrical resistance was measured for 1 h at each concentration.

3. Results and Discussion

3.1. Role of Reaction Temperature and Duration on the Formation of MoO3

Based on their XRD patterns, MoO3-1 and MoO3-2 were identified as pure h-MoO3 (ICDD 065-0033, Figure 1a,b). Every reflection belonged to this phase indicating that no contaminations were present. MoO3-3, which were prepared at 210 °C for 3 h, was composed of the h- and o-MoO3 phase (ICDD 01-074-7909); however, the longer reaction duration, namely 6 h, resulted in pure o-MoO3 (Figure 1c,d). As the temperature was further increased to 240 °C, at both reaction times pure o-MoO3 was obtained (Figure 1e,f). All samples were crystalline, demonstrated by the sharp, intensive peaks.
According to SEM images, the h-MoO3 phases prepared at 90 °C were crystallized in a well-distinguished hexagonal form at both reaction durations (Figure 2, MoO3-1-2). The rods had a diameter of 7–12 μm when a 3 h duration was applied, and 8–18 μm diameter with a 6 h reaction time; however, their length was 10–50 μm in both cases. It can be also observed that the MoO3-2 sample had a much smoother surface and sharp edges due to the longer reaction time.
In the case of MoO3-3, the sample contained both h- and o-MoO3 phases, whose presence was obvious based on the observed morphologies. In addition to the robust hexagonal rods, 200–500 nm thick fibers formed, with longer (30–50 µm) lengths (in some cases they were broken into shorter, i.e., 1–2 µm pieces), referring to the o-MoO3 phase (Figure 2, MoO3-3). The morphology of MoO3-4 was, however, homogenous, and the sample was composed of 250–800 nm thick and 10–50 µm long fibers owing to the presence of only the o-MoO3 phase (Figure 2, MoO3-4).
When increasing the temperature to 240 °C, both reaction durations resulted in o-MoO3 with fibrous morphology. The diameter of the fibers was 250–800 nm when 3 h was applied, but the fibers became 500–1000 nm thick with 6 h reaction time. Their length was several 10–50 μm in both cases (Figure 2, MoO3-5-6).
Based on these results, it was found that the crystalline structure and morphology of the obtained products can be easily influenced by the appropriate choice of reaction temperature and duration. The lower temperature of 90 °C resulted in a pure hexagonal MoO3 structure with the characteristic hexagonal micrometer thick rod-like morphology, whereas the higher temperature, namely 240 °C led to a pure orthorhombic MoO3 phase, which was crystallized in the form of nanofibers. The transformation of the hexagonal to orthorhombic phase could be observed using 210 °C, and 3 h synthesis parameters, where both polymorphs were present with both types of morphologies. However, not only the reaction temperature but also the duration played an important role in this phase transformation, as it took place completely only when a long enough reaction time (6 h) was applied.

3.2. Role of Additives on the Formation of MoO3

When the CTAB additive was used, samples MoO3-7 and also MoO3-8 were identified as pure crystalline o-MoO3, similarly to those reactions, where no additive was used (Figure 3a,b). Experiments with the addition of CrCl3, however, resulted in a pure, metastable m-MoO3 phase (ICDD 01-085-2405) at both 3 and 6 h reaction durations, which is unprecedented in the literature (Figure 3c,d).
In the case of CTAB additive, the uniform morphology typical of o-MoO3 formed here as well, containing fibers with a diameter of 200–500 nm and a length of 10–30 µm, which were slightly fragmented in the case of 6 h reaction time (Figure 4, MoO3-7-8).
The new metastable m-MoO3 was crystallized in the form of nanosheets, in contrast with the former, longitudinal shapes (columns, rods) of the other phases. The sheets had an average thickness of 100–250 nm when 3 h reaction time was applied and it increased to 100–500 nm at 6 h duration, due to the longer reaction time which favored crystal growth. The sheets were 10–20 μm long in both cases and were arranged almost parallel to each other (Figure 4, MoO3-9-10).
Examining the role of the additives, it was concluded that the CTAB additive did not influence the formation of crystalline phases during the reaction, as it also led to pure o-MoO3. On the other hand, CTAB affected the morphology resulting in a significant reduction in the diameter of the fibers. Without the additive the fiber diameter of o-MoO3 was 250–800 nm and 500–1000 nm, depending on the duration (3 h and 6 h, respectively), but with CTAB it became much thinner, 100–250 nm and 100–500 nm in the 3 and the 6 h experiments, respectively. In contrast, the addition of CrCl3 additive resulted in the formation of the metastable m-MoO3, which has not been prepared by hydrothermal synthesis previously yet. Its role in directing crystal growth also became obvious as it led to nanosheet morphology, presumably due to its favored adherence to certain crystal planes resulting in inhibition of crystal growth in specified directions.

3.3. Further Investigation of the Different MoO3 Phases

For a more detailed examination of hexagonal (MoO3-1), o- (MoO3-8), and m-MoO3 (MoO3-10) phases, various methods were used.
Results of the specific surface area measurement demonstrated significant differences among the phases, which can be explained by their completely different morphologies (Table 2 SBET). The smallest (0.21 m2/g) specific surface area belonged to the hexagonal MoO3 phase, which was crystallized in the form of the most robust hexagonal rods with a diameter and length of several µm. The orthorhombic phase had the largest (9.7 m2/g) surface area, which is attributed to its nanofibrous morphology, whereas the size of monoclinic MoO3 nanosheets surface was 2.9 m2/g.
Studying the elemental composition of the phases by EDX, it was found that the ratio of the main elements (Mo, O) was close to the stoichiometric Mo:O = 1:3 ratio, and there were no other elements referring to contaminations in the samples.
In the FT-IR spectra, the characteristic metal-oxygen (Mo-O) vibrations in the range below 1000 cm−1 can be observed, by the help of which the phases can be distinguished (Figure 5A). Since the orthorhombic and monoclinic crystal structures are very similar, their spectrum is hardly different, unlike the hexagonal phase, where the difference is clearly visible. The intensive peak around 1000–900 cm−1 refers to the Mo=O band, while the other one at 550–600 cm−1 belongs to the Mo-O valence vibration. In the case of o-MoO3, another less intensive peak appears at 830 cm−1, which is attributed to the Mo-O-Mo deformation vibration and it is not present in the spectrum of the monoclinic phase. In contrast, in the spectrum of the h-MoO3 phase, a band can be observed at 700–750 cm−1 referring to the Mo-O vibration, which does not appear in the case of the other two phases. In addition, peaks around 1400 and 3200 cm−1, moreover at 1600 and 3400 cm−1 belong to the N-H vibrations of the NH4+ group and to the O-H vibrations of O-H groups and water molecules, respectively. These bands are not detectable in the case of orthorhombic and monoclinic MoO3 phases. This phenomenon can be explained by the fact that hexagonal channels usually store small ions, e.g., NH4+ or molecules e.g., H2O which stabilize the structure, similar to h-WO3 [2,11,15,34,35].
In the Raman spectra, the distinctive bands of o-MoO3 at 990, 820, and 680 cm−1 belong to the stretching mode of Mo=O1, Mo-O3-Mo, and the Mo-O2-Mo (Figure 5B). The peaks appearing between 400 and 170 cm−1 refer to the different deformation vibrations of O=Mo=O and O-Mo-O, while those below 170 cm−1 are derived from the translation of rigid MoO4 chain. The monoclinic phase has a similar Raman spectrum, but at 500 cm−1 there is a band of the Mo-O valence vibration that is absent in the spectrum of the orthorhombic MoO3 and there are evident differences between o- and m-MoO3 in the ratio of the peaks, as well, mostly in the range of deformation vibrations. In the case of h-MoO3, the band at 1000 cm−1 and the triple peak located at 910–900–880 cm−1 are ascribed to the Mo=O valence vibrations, whereas the bands around 700, 500, 390, and 320 cm−1 refer to the O-Mo-O vibrations. The peaks below 300 cm−1 are attributed to the modes of the MoO4 tetrahedra chains [36,37,38,39].
The absorption properties of the phases were investigated by diffuse reflectance UV–Vis spectroscopy. In each spectrum, a distinct absorption edge can be observed, which is characteristic of semiconductors (Figure 5C). Every phase absorbs not only in the UV range, but also in the visible part of the spectrum. The hexagonal phase, due to its pale yellow color, absorbs only below 500 nm, while the blue colored orthorhombic phase has absorption also above 550 nm, and so does monoclinic MoO3. Based on the spectra, the band gaps were determined, which were 3.07 eV for h-MoO3, 3.33 eV for o-MoO3, and 3.02 eV for m-MoO3.

3.4. Gas Sensing

3.4.1. Checking Crystalline Phases and Morphology at 500 °C

The gas sensing properties were tested between room temperature up to 500 °C. However, at lower temperatures the gas sensing signals were so small that they could not be evaluated properly. Therefore, the responses to various H2 gas concentrations were only recorded at 500 °C. Accordingly, the present study serves only as a short test about the gas sensing of hydrothermally prepared m-MO3, but later a more detailed study is needed with lower sensing temperatures and various test gases.
Due to the relatively high operating temperature, before the gas sensing investigations, it was needed to check the possible phase and morphology transformations occurring at 500 °C, at the temperature of the gas sensing tests. Hence, all the MoO3 polymorphs had a 1 h heat treatment at 500 °C, and then they were studied by XRD and SEM (Figure 6).
Based on SEM and XRD results, we concluded that the heat treatment did not change the phase and morphology of o- and m-MoO3. However, the initially h-MoO3 phase transformed into o-MoO3, and the morphology changed, as well. The heated h-MoO3 (now o-MoO3) samples consisted of no hexagonal rods as before, but 10–20 nm wide, 200–300 nm long side-adhered plates, and irregular particles in various sizes (10–500 nm, Figure 6).

3.4.2. Gas Sensing Tests

During the tests the samples gave an n-type response, namely, their electrical resistance was reduced when exposed to hydrogen, which is typical for an n-type semiconductor reacting to a reducing gas [40]. The change in the logarithm of their electrical resistance as a function of time is presented in (Figure 7A–C). The sensitivity (S) of the sensors can be calculated using the following formula [41]: S = | Δ R | R b a s e , where ∆R is the change in electrical resistance due to the presence of the gas, while Rbase is the baseline electrical resistance of the material. As the baseline did not slip significantly, we selected the measured electrical resistance before the introduction of the 25,000 ppm H2 gas as Rbase. It can be observed that the sensitivity does not increase linearly with the increasing H2 concentration, and this might be related to the saturation of the gas sensing active sites (Figure 7D). Based on the gas sensing results, m- (β-) MoO3 is more sensitive to the H2 gas than o- (α-) MoO3 which has, however, a larger specific surface area. Probably the density of surface active site is different in their case, which can be explained by, e.g., different amounts and forms of pre-adsorbed oxygen (O2−, O or O2) or the difference in surface OH groups similar to m- and h-WO3 [40,42]. The annealed h-MO3, which transformed into o-MoO3 at 500 °C has a similar sensitivity to the monoclinic polymorph. Probably the hexagonal-orthorhombic transformation caused an increase in the number of active sites or in the surface area, which might explain this.
The response time was defined as the time at which the electrical resistance reaches 90% of the maximum signal given to a certain gas concentration and the recovery time was defined as the time needed to return to 10% of the maximum signal after finishing the gas pulse. These values are given for each sample using 25,000 ppm H2 (Table 3). Based on the data, the response time of all samples is low, approx. 1 min. The samples with a similar sensitivity (monoclinic and transformed hexagonal) also have a similar recovery time, approx. 20 min. The recovery time was lower for the fiber-like o-MoO3 sample. This might be explained by its lower number of gas sensing active sites, which also explained the lower sensitivity of this sample.

4. Conclusions

In this study, we investigated the effect of various parameters on the hydrothermal preparation of MoO3, for the first time also at 240 °C, focusing on the changes in the crystalline phases and morphology of the products. We used different temperatures (90/210/240 °C) and durations (3/6 h) to investigate their role in our reaction system. At 240 °C, we also tested the effect of CTAB and CrCl3 additives, respectively. As a result, h-MoO3 formed at 90 °C independently on the duration and crystallized in the form of hexagonal rods. At 210 °C with 3 h reaction time, the product was a mixture of h- and o-MoO3, while with 6 h it was pure o-MoO3, alike at 240 °C, with nanofibrous morphology. With the use of the CTAB additive, we obtained pure o-MoO3 in the form of nanofibers, but with a radically reduced diameter due to the presence of the additive. However, when we applied CrCl3, it resulted in the metastable m-MoO3 phase at both reaction durations with nanosheet morphology. The gas sensing properties of the MoO3 samples against H2 were also tested. The m-MoO3 was found to be more sensitive than the o-MoO3.

Author Contributions

Conceptualization, T.N.-K., I.M.S., and G.P.; methodology, T.N.-K. and I.M.S.; investigation, T.N.-K., L.S., I.E.L., K.L., and P.P.; writing—original draft preparation, T.N.-K.; writing—review and editing, K.L., P.P., I.M.S., and G.P.; supervision, I.M.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by NRDI grant numbers K 124212, TNN_16 123631. The work performed within project no. VEKOP-2.3.2-16-2017-00013 was supported by the European Union and the State of Hungary, co-financed by the European Regional Development Fund. The research reported in this paper was supported by the Higher Education Excellence Program of the Ministry of Human Capacities in the frame of Nanotechnology and Materials Science research area of Budapest University of Technology (BME FIKP-NAT). The APC was funded by the Budapest University of Technology and Economics.

Acknowledgments

The authors thank T. Igricz (Budapest University of Technology and Economics, Department of Organic Chemistry and Technology) for help in the Raman measurements.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Krobkrong, N.; Itthibenchapong, V.; Khongpracha, P.; Faungnawakij, K. Deoxygenation of oleic acid under an inert atmosphere using molybdenum oxide-based catalysts. Energy Convers. Manag. 2018, 167, 1–8. [Google Scholar] [CrossRef]
  2. Alam, U.; Kumar, S.; Bahnemann, D.; Koch, J.; Tegenkamp, C.; Muneer, M. Harvesting visible light with MoO3 nanorods modified by Fe(iii) nanoclusters for effective photocatalytic degradation of organic pollutants. Phys. Chem. Chem. Phys. 2018, 20, 4538–4545. [Google Scholar] [CrossRef] [PubMed]
  3. Yogananda, H.S.; Nagabhushana, H.; Darshan, G.P.; Basavaraj, R.B. MoO3 nanostructures from EGCG assisted sonochemical route: Evaluation of its application towards forensic and photocatalysis. J. Alloys Compd. 2018, 745, 874–891. [Google Scholar] [CrossRef]
  4. Li, B.; Wang, X.; Wu, X.; He, G.; Xu, R.; Lu, X.; Wang, F.R.; Parkin, I.P. Phase and morphological control of MoO3−x nanostructure for efficient cancer theragnosis therapy. Nanoscale 2017, 9, 11012–11016. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Song, Y.; Zhao, J.; Zhao, Y.; Huang, Z.; Li, Y.; Wu, G. Synthesis of α-MoO3 nanobelts with preferred orientation and good photochromic performance. CrystEngComm 2016, 18, 6502–6512. [Google Scholar] [CrossRef]
  6. Xie, Z.; Feng, Y.; Wang, F.; Chen, D.; Zhang, Q.; Zeng, Y.; Lv, W.; Liu, G. Construction of carbon dots modified MoO3/g-C3N4Z-scheme photocatalyst with enhanced visible-light photocatalytic activity for the degradation of tetracycline. Appl. Catal. B Environ. 2018, 229, 96–104. [Google Scholar] [CrossRef]
  7. Phuruangrat, A.; Cheed-im, U.; Thongtem, T.; Thongtem, S. Influence of Gd dopant on photocatalytic properties of MoO3 nanobelts. Mater. Lett. 2016, 173, 158–161. [Google Scholar] [CrossRef]
  8. Chithambararaj, A.; Sanjini, N.S.; Velmathi, S.; Bose, A.C. Preparation of h-MoO3 and a-MoO3 nanocrystals: Comparative study on photocatalytic degradation of methylene blue under visible light irradiation. Phys. Chem. Chem. Phys. 2013, 15, 14761–14769. [Google Scholar] [CrossRef]
  9. Illyaskutty, N.; Kohler, H.; Trautmann, T.; Schwotzer, M.; Mahadevan Pillai, V.P. Hydrogen and ethanol sensing properties of molybdenum oxide nanorods based thin films: Effect of electrode metallization and humid ambience. Sens. Actuators B Chem. 2013, 187, 611–621. [Google Scholar] [CrossRef]
  10. Liu, Y.; Yang, S.; Lu, Y.; Podval’naya, N.V.; Chen, W.; Zakharova, G.S. Hydrothermal synthesis of h-MoO3 microrods and their gas sensing properties to ethanol. Appl. Surf. Sci. 2015, 359, 114–119. [Google Scholar] [CrossRef]
  11. Xu, Y.; Xie, L.; Zhang, Y.; Cao, X. Hydrothermal synthesis of hexagonal MoO3 and its reversible electrochemical behavior as a cathode for Li-ion batteries. Electron. Mater. Lett. 2013, 9, 693–696. [Google Scholar] [CrossRef]
  12. Han, B.; Lee, K.; Lee, Y.; Kim, S.; Park, H.; Hwang, B.; Kwak, D.; Park, K. MoO3 nanostructured electrodes prepared via hydrothermal process for lithium ion batteries. Int. J. Electrochem. Sci. 2015, 10, 4232–4240. [Google Scholar]
  13. Dhanasankar, M.; Purushothaman, K.K.; Muralidharan, G. Optical, structural and electrochromic studies of molybdenum oxide thin films with nanorod structure. Solid State Sci. 2010, 12, 246–251. [Google Scholar] [CrossRef]
  14. Chithambararaj, A.; Rajeswari Yogamalar, N.; Bose, A.C. Hydrothermally synthesized h-MoO3 and α-MoO3 nanocrystals: New findings on crystal-structure-dependent charge transport. Cryst. Growth Des. 2016, 16, 1984–1995. [Google Scholar] [CrossRef]
  15. Kumar, V.V.; Gayathri, K.; Anthony, S.P. Synthesis of α-MoO3 nanoplates using organic aliphatic acids and investigation of sunlight enhanced photodegradation of organic dyes. Mater. Res. Bull. 2016, 76, 147–154. [Google Scholar] [CrossRef]
  16. 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. [Google Scholar] [CrossRef]
  17. Szilágyi, I.M.; Madarász, J.; Pokol, G.; Király, P.; Tárkányi, G.; Saukko, S.; Mizsei, J.; Tóth, A.L.; Szabó, A.; Varga-Josepovits, K. Stability and controlled composition of hexagonal WO3. Chem. Mater. 2008, 20, 4116–4125. [Google Scholar] [CrossRef]
  18. Badr, A.M.; El-Anssary, E.H.; Elshaikh, H.A.; Afify, H.H. Hydrothermal synthesis and influence of later heat treatment on the structural evolution, optical and electrical properties of nanostructured α-MoO3 single crystals. J. Phys. D Appl. Phys. 2017, 50, 505111. [Google Scholar] [CrossRef]
  19. Bouzidi, A.; Benramdane, N.; Tabet-derraz, H.; Mathieu, C.; Khelifa, B.; Desfeux, R. Effect of substrate temperature on the structural and optical properties of MoO3 thin films prepared by spray pyrolysis technique. Mater. Sci. Eng. B 2003, 97, 5–8. [Google Scholar] [CrossRef]
  20. 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]
  21. Yan, H.; Song, P.; Zhang, S.; Zhang, J.; Yang, Z.; Wang, Q. Au nanoparticles modified MoO3 nanosheets with their enhanced properties for gas sensing. Sens. Actuators B Chem. 2016, 236, 201–207. [Google Scholar] [CrossRef]
  22. Chang, W.C.; Qi, X.; Kuo, J.C.; Lee, S.C.; Ng, S.K.; Chen, D. Post-deposition annealing control of phase and texture for the sputtered MoO3 films. CrystEngComm 2011, 13, 5125–5132. [Google Scholar] [CrossRef]
  23. Haro-Poniatowski, E.; Jouanne, M.; Morhange, J.F.; Julien, C.; Diamant, R.; Fernández-Guasti, M.; Fuentes, G.A.; Alonso, J.C. Micro-Raman characterization of WO3 and MoO3 thin films obtained by pulsed laser irradiation. Appl. Surf. Sci. 1998, 127–129, 674–678. [Google Scholar] [CrossRef]
  24. Maiti, P.; Guha, P.; Singh, R.; Dash, J.K.; Satyam, P.V. Optical band gap, local work function and field emission properties of MBE grown B-MoO3 nanoribbons. Appl. Surf. Sci. 2019, 476, 691–700. [Google Scholar] [CrossRef]
  25. Wang, L.; Zhang, G.-H.; Sun, Y.-J.; Zhou, X.-W.; Chou, K.-C. Preparation of ultrafine β-MoO3 from industrial grade MoO3 powder by the method of sublimation. J. Phys. Chem. C 2016, 120, 19821–19829. [Google Scholar] [CrossRef]
  26. Mizushima, T.; Moriya, Y.; Phuc, N.H.H.; Ohkita, H.; Kakuta, N. Soft chemical transformation of α-MoO3 to β-MoO3 as a catalyst for vapor-phase oxidation of methanol. Catal. Commun. 2011, 13, 10–13. [Google Scholar] [CrossRef]
  27. Pham, T.T.P.; Nguyen, P.H.D.; Vo, T.T.; Nguyen, H.H.P.; Luu, C.L. Facile method for synthesis of nanosized β–MoO3 and their catalytic behavior for selective oxidation of methanol to formaldehyde. Adv. Nat. Sci. Nanosci. Nanotechnol. 2015, 6, 045010. [Google Scholar] [CrossRef]
  28. Alsaif, M.M.Y.A.; Balendhran, S.; Field, M.R.; Latham, K.; Wlodarski, W.; Ou, J.Z.; Kalantar-Zadeh, K. Two dimensional α-MoO3 nanoflakes obtained using solvent-assisted grinding and sonication method: Application for H2 gas sensing. Sens. Actuators B Chem. 2014, 192, 196–204. [Google Scholar] [CrossRef]
  29. Arachchige, H.M.M.M.; Comini, E.; Zappa, D.; Sberveglieri, G. Gas Sensing Properties of MoO3. Proceedings 2017, 1, 449. [Google Scholar] [CrossRef] [Green Version]
  30. Mane, A.A.; Moholkar, A.V. Orthorhombic MoO3 nanobelts based NO2 gas sensor. Appl. Surf. Sci. 2017, 405, 427–440. [Google Scholar] [CrossRef]
  31. Martínez, H.M.; Torres, J.; Rodríguez-García, M.E.; López Carreño, L.D. Gas sensing properties of nanostructured MoO3 thin films prepared by spray pyrolysis. Phys. B Condens. Matter 2012, 407, 3199–3202. [Google Scholar] [CrossRef]
  32. Brunauer, S.; Emmett, P.H.; Teller, E. Adsorption of gases in multimolecular layers. J. Am. Chem. Soc. 1938, 60, 309. [Google Scholar] [CrossRef]
  33. Szilágyi, I.M.; Bakos, L.P.; Nóra, J.; Moura da Silva Bezerra da Costa, U.C.; László, K.; Lábár, J.; Igricz, T.; Varga-Josepovits, K.; Pasierb, P.; Färm, E.; et al. Photocatalytic and gas sensitive multiwalled carbon prepared by atomic layer deposition. Nanomaterials 2020, 10, 252. [Google Scholar]
  34. Liu, Y.; Feng, P.; Wang, Z.; Jiao, X.; Akhtar, F. Novel fabrication and enhanced photocatalytic MB degradation of hierarchical porous monoliths of MoO3 nanoplates. Sci. Rep. 2017, 7, 1845. [Google Scholar] [CrossRef]
  35. Chithambararaj, A.; Bose, A.C. Hydrothermal synthesis of hexagonal and orthorhombic MoO3 nanoparticles. J. Alloys Compd. 2011, 509, 8105–8110. [Google Scholar] [CrossRef]
  36. Sinaim, H.; Jin, D.; Sung, J.; Phuruangrat, A.; Thongtem, S. Free-polymer controlling morphology of L-MoO3 nanobelts by a facile hydrothermal synthesis, their electrochemistry for hydrogen evolution reactions and optical properties. J. Alloys Compd. 2012, 516, 172–178. [Google Scholar] [CrossRef]
  37. Gao, B.; Fan, H.; Zhang, X. Hydrothermal synthesis of single crystal MoO3 nanobelts and their electrochemical properties as cathode electrode materials for rechargeable lithium batteries. J. Phys. Chem. Solids 2012, 73, 423–429. [Google Scholar] [CrossRef]
  38. Senthilkumar, R.; Anandhababu, G.; Mahalingam, T.; Ravi, G. Photoelectrochemical study of MoO3 assorted morphology films formed by thermal evaporation. J. Energy Chem. 2016, 25, 798–804. [Google Scholar] [CrossRef]
  39. Hui, B.; Li, G.; Zhao, X.; Wang, L.; Wu, D.; Li, J.; Via, B.K. h-MoO3 microrods grown on wood substrates through a low-temperature hydrothermal route and their optical properties. J. Mater. Sci. Mater. Electron. 2017, 28, 3264–3271. [Google Scholar] [CrossRef]
  40. Szilágyi, I.M.; Saukko, S.; Mizsei, J.; Tóth, A.L.; Madarász, J.; Pokol, G. Gas sensing selectivity of hexagonal and monoclinic WO3 to H2S. Solid State Sci. 2010, 12, 1857–1860. [Google Scholar] [CrossRef]
  41. Chaudhari, G.N.; Bende, A.M.; Bodade, A.B.; Patil, S.S.; Sapkal, V.S. Structural and gas sensing properties of nanocrystalline TiO2:WO3-based hydrogen sensors. Sens. Actuators B Chem. 2006, 115, 297–302. [Google Scholar] [CrossRef]
  42. Szilágyi, I.M.; Fórizs, B.; Rosseler, O.; Szegedi, Á.; Németh, P.; Király, P.; Tárkányi, G.; Vajna, B.; Varga-Josepovits, K.; László, K.; et al. WO3 photocatalysts: Influence of structure and composition. J. Catal. 2012, 294, 119–127. [Google Scholar] [CrossRef] [Green Version]
Figure 1. XRD patterns of the samples synthetized in the hydrothermal reaction of AHM and HNO3 at different temperatures and durations, (a) MoO3-1 (90 °C, 3 h), (b) MoO3-2 (90 °C, 6 h), (c) MoO3-3 (210 °C, 3 h), (d) MoO3-4 (210 °C, 6 h), (e) MoO3-5 (240 °C, 3 h), (f) MoO3-6 (90 °C, 3 h).
Figure 1. XRD patterns of the samples synthetized in the hydrothermal reaction of AHM and HNO3 at different temperatures and durations, (a) MoO3-1 (90 °C, 3 h), (b) MoO3-2 (90 °C, 6 h), (c) MoO3-3 (210 °C, 3 h), (d) MoO3-4 (210 °C, 6 h), (e) MoO3-5 (240 °C, 3 h), (f) MoO3-6 (90 °C, 3 h).
Nanomaterials 10 00891 g001
Figure 2. SEM images of the samples synthetized in the hydrothermal reaction of AHM and HNO3 at different temperatures and durations, MoO3-1 90 °C, 3 h, MoO3-2: 90 °C, 6 h, MoO3-3: 210 °C, 3 h, MoO3-4: 210 °C, 6 h, MoO3-5: 240 °C, 3 h, MoO3-6: 240 °C, 6 h.
Figure 2. SEM images of the samples synthetized in the hydrothermal reaction of AHM and HNO3 at different temperatures and durations, MoO3-1 90 °C, 3 h, MoO3-2: 90 °C, 6 h, MoO3-3: 210 °C, 3 h, MoO3-4: 210 °C, 6 h, MoO3-5: 240 °C, 3 h, MoO3-6: 240 °C, 6 h.
Nanomaterials 10 00891 g002
Figure 3. XRD patterns of the samples synthetized in the hydrothermal reaction of AHM and HNO3 with various additives, (a) MoO3-7 (CTAB, 240 °C, 3 h), (b) MoO3-8 (CTAB, 240 °C, 6 h), (c) MoO3-9 (CrCl3, 240 °C, 3 h), (d) MoO3-10 (CrCl3, 240 °C, 6 h).
Figure 3. XRD patterns of the samples synthetized in the hydrothermal reaction of AHM and HNO3 with various additives, (a) MoO3-7 (CTAB, 240 °C, 3 h), (b) MoO3-8 (CTAB, 240 °C, 6 h), (c) MoO3-9 (CrCl3, 240 °C, 3 h), (d) MoO3-10 (CrCl3, 240 °C, 6 h).
Nanomaterials 10 00891 g003
Figure 4. SEM images of the samples synthetized in the hydrothermal reaction of AHM and HNO3 with various additives, MoO3-7: CTAB, 240 °C, 3 h, MoO3-8: CTAB, 240 °C, 6 h, MoO3-9: CrCl3, 240 °C, 3 h, MoO3-10: CrCl3, 240 °C, 6.
Figure 4. SEM images of the samples synthetized in the hydrothermal reaction of AHM and HNO3 with various additives, MoO3-7: CTAB, 240 °C, 3 h, MoO3-8: CTAB, 240 °C, 6 h, MoO3-9: CrCl3, 240 °C, 3 h, MoO3-10: CrCl3, 240 °C, 6.
Nanomaterials 10 00891 g004
Figure 5. FT-IR (A), Raman (B), and UV–Vis (C) spectra of different MoO3 phases, (a) MoO3-10, monoclinic phase (CrCl3, 240 °C, 6 h), (b) MoO3-8, orthorhombic phase (CTAB, 240 °C, 6 h), (c) MoO3-1, hexagonal phase (90 °C, 3 h).
Figure 5. FT-IR (A), Raman (B), and UV–Vis (C) spectra of different MoO3 phases, (a) MoO3-10, monoclinic phase (CrCl3, 240 °C, 6 h), (b) MoO3-8, orthorhombic phase (CTAB, 240 °C, 6 h), (c) MoO3-1, hexagonal phase (90 °C, 3 h).
Nanomaterials 10 00891 g005aNanomaterials 10 00891 g005b
Figure 6. XRD patterns (A) and SEM images of different MoO3 phases after the 500 °C heat treatment., (a) MoO3-10, monoclinic phase (CrCl3, 240 °C, 6 h), (b) MoO3-8, orthorhombic phase (CTAB, 240 °C, 6 h), (c) MoO3-1, hexagonal phase (90 °C, 3 h).
Figure 6. XRD patterns (A) and SEM images of different MoO3 phases after the 500 °C heat treatment., (a) MoO3-10, monoclinic phase (CrCl3, 240 °C, 6 h), (b) MoO3-8, orthorhombic phase (CTAB, 240 °C, 6 h), (c) MoO3-1, hexagonal phase (90 °C, 3 h).
Nanomaterials 10 00891 g006
Figure 7. Results of the gas sensing test of the different MoO3 phases. (A) MoO3-10, monoclinic phase (CrCl3, 240 °C, 6 h); (B) MoO3-8, orthorhombic phase (CTAB, 240 °C, 6 h); (C) MoO3-1, hexagonal phase (90 °C, 3 h); (D) sensitivity of the MoO3 polymorphs.
Figure 7. Results of the gas sensing test of the different MoO3 phases. (A) MoO3-10, monoclinic phase (CrCl3, 240 °C, 6 h); (B) MoO3-8, orthorhombic phase (CTAB, 240 °C, 6 h); (C) MoO3-1, hexagonal phase (90 °C, 3 h); (D) sensitivity of the MoO3 polymorphs.
Nanomaterials 10 00891 g007
Table 1. Experimental conditions of the hydrothermal reaction between AHM and HNO3.
Table 1. Experimental conditions of the hydrothermal reaction between AHM and HNO3.
SampleAdditiveT (°C)t (h)
MoO3-1-903
MoO3-2-906
MoO3-3-2103
MoO3-4-2106
MoO3-5-2403
MoO3-6-2406
MoO3-7CTAB2403
MoO3-8CTAB2406
MoO3-9CrCl32403
MoO3-10CrCl32406
Table 2. The specific surface area (SBET) and EDX results of the different MoO3 phases, MoO3-1: 90 °C, 3 h, MoO3-8: CTAB, 240 °C, 6 h, MoO3-11: CrCl3, 240 °C, 6 h.
Table 2. The specific surface area (SBET) and EDX results of the different MoO3 phases, MoO3-1: 90 °C, 3 h, MoO3-8: CTAB, 240 °C, 6 h, MoO3-11: CrCl3, 240 °C, 6 h.
SampleSBET (m2/g)EDX (Atom %)
Mo (Average)O (Average)
MoO3-1, hexagonal0.2124.375.7
MoO3-8, orthorhombic9.724.875.2
MoO3-10, monoclinic2.925.774.3
Table 3. Response and return times of sensors using 25,000 ppm H2.
Table 3. Response and return times of sensors using 25,000 ppm H2.
SampleResponse Time (min)Recovery Time (min)
MoO3-10, monoclinic1.222.5
MoO3-8, orthorhombic1.210.1
MoO3-1, hexagonal1.220.3

Share and Cite

MDPI and ACS Style

Nagyné-Kovács, T.; Studnicka, L.; Lukács, I.E.; László, K.; Pasierb, P.; Szilágyi, I.M.; Pokol, G. Hydrothermal Synthesis and Gas Sensing of Monoclinic MoO3 Nanosheets. Nanomaterials 2020, 10, 891. https://doi.org/10.3390/nano10050891

AMA Style

Nagyné-Kovács T, Studnicka L, Lukács IE, László K, Pasierb P, Szilágyi IM, Pokol G. Hydrothermal Synthesis and Gas Sensing of Monoclinic MoO3 Nanosheets. Nanomaterials. 2020; 10(5):891. https://doi.org/10.3390/nano10050891

Chicago/Turabian Style

Nagyné-Kovács, Teodóra, Levente Studnicka, István Endre Lukács, Krisztina László, Pawel Pasierb, Imre Miklós Szilágyi, and György Pokol. 2020. "Hydrothermal Synthesis and Gas Sensing of Monoclinic MoO3 Nanosheets" Nanomaterials 10, no. 5: 891. https://doi.org/10.3390/nano10050891

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