Next Article in Journal
Adsorbent Material Based on Carbon Black and Bismuth with Tunable Properties for Gold Recovery
Previous Article in Journal
Recovery of Valuable Materials from End-of-Life Photovoltaic Solar Panels
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Thermal Stability, Optical and Electrical Properties of Substoichiometric Molybdenum Oxide

1
School of Materials Science and Engineering, Zhengzhou University, Zhengzhou 450001, China
2
Zhongyuan Critical Metals Laboratory, Zhengzhou University, Zhengzhou 450001, China
3
Central China Branch, Oriental Green Energy (Hebei) Co., Ltd., Zhengzhou 450003, China
*
Authors to whom correspondence should be addressed.
Materials 2023, 16(7), 2841; https://doi.org/10.3390/ma16072841
Submission received: 20 February 2023 / Revised: 19 March 2023 / Accepted: 21 March 2023 / Published: 2 April 2023
(This article belongs to the Section Advanced and Functional Ceramics and Glasses)

Abstract

:
Substoichiometric molybdenum oxide ceramics have aroused widespread interest owing to their promising optical and electrical performance. In this work, the thermal stability and decomposition mechanism of Mo9O26 and Mo4O11 at 700–1000 °C and 700–1100 °C were investigated, respectively. Based on this information, MoOx (2 < x < 3) bulk ceramics were prepared by spark plasma sintering (SPS). The results show that Mo9O26 is stable up to 790 °C in an argon atmosphere. As the temperature rises, it decomposes into Mo4O11. Mo4O11 can exist stably at 830 °C, beyond which it will convert to MoO2. The MoOx ceramic bulks with four different components (MoO2.9, MoO2.8, MoO2.7 and MoO2.6) were successfully sintered by SPS, and their relative density was greater than 96.4% as measured by the Archimedes principle. The reflectivity of MoOx ceramic bulk is low and only 6.3% when the composition is MoO2.8. The resistivity increases from 10−3 to 10−1 Ωcm with the increase in the O/Mo atomic ratio x. In general, the thermal stability information provides a theoretical basis for the processing of MoOx materials, such as the sintering of the MoOx target. The optical and electrical properties show that MoOx is a low-reflective conductive oxide material with great photoelectric application value.

1. Introduction

Molybdenum oxide comes in various forms, including molybdenum trioxide (MoO3) with a wide gap (>2.7 eV), substoichiometric molybdenum oxide (MoOx) with oxygen vacancies and semi-metallic molybdenum dioxide with metallic properties. Due to its unique physical and chemical properties, molybdenum trioxide has been applied in many fields, such as photocatalysis [1,2,3], hole injector for organic solar cells [4,5,6] and high-capacity lithium-ion batteries [7,8,9]. Among the applications of molybdenum oxide, substoichiometric molybdenum oxide has gradually received more attention because of its excellent optical, electrical and catalytic properties. In the field of optoelectronics, MoOx can be used in the antireflection layer of displays [10,11], to improve solar energy conversion efficiency [12,13,14,15] and in new-generation sodium magnesium batteries [16,17,18]. In addition, it can also be used in catalysis [19,20], biomedicine [21], electronic devices [22,23] and other fields.
Although MoOx has excellent properties, little research has been conducted on its thermal stability. As weakly stable oxides, Gunnar Hägg and Arne Magnéli found that Mo4O11 decomposes at 700–850 °C, and Mo9O26 converts to Mo8O23 at 765 °C; it has also been proposed that Mo4O11 melts into a liquid phase and MoO2 at 818 °C, and Mo9O26 melts into Mo4O11 and a liquid phase at 780 °C [24]. In recent years, studies on its stability have mainly focused on the usage scenarios of two-dimensional (2D) materials. Jinyoun Cho et al. [25] and Shuangying Cao et al. [26] found that the maximum heat treatment temperature of the MoOx layer in silicon heterojunction solar cells should not exceed 170–200 °C, otherwise a significant redox reaction and metal atom diffusion will occur. The study also found that annealing under argon can effectively improve the thermal stability of MoOx [27,28]. It can also be stacked with high work function metals, e.g., Ni, and subjected to certain heat treatments for higher stability [29]. Hennrik Schmidt et al. [11] found that the original amorphous MoOx films prepared by magnetron sputtering began to transform into crystalline MoO2 films after thermal treatment at 350 °C. It can be observed that this substoichiometric molybdenum oxide exists stably only within a specific temperature range, and the thermal stability, decomposition mechanism and phase generated in the decomposition process have not been studied in detail. Therefore, it is necessary to explore its thermal stability and provide a theoretical basis for expanding its application range, such as preparing MoOx ceramic targets, thus producing films with a specific O/Mo ratio.
In this work, the Mo9O26 and Mo4O11 substoichiometric molybdenum oxide powders were heated in argon to explore the thermal stability and decomposition behavior of these two powders, and the sintering process was determined according to the thermal stability. Four ceramic bulks, MoO2.9, MoO2.8, MoO2.7 and MoO2.6, were prepared by SPS, and the relative density, reflectivity and resistivity were characterized. This work provides useful information for the use of substoichiometric molybdenum oxide in optical and electrical fields.

2. Materials and Methods

2.1. Preparation of Substoichiometric Molybdenum Oxide Powders

Two powders with main phases of Mo4O11 (7.47 µm) and Mo9O26 (5.58 µm) are prepared. The MoO3 (99.95%, 2.89 µm) powder is placed into a corundum crucible (80 × 40 × 20 mm3) and then into a tubular furnace. The tube furnace is evacuated to vacuum (−0.1 MPa); then, argon is injected at a flow rate of 70 mL/min. The MoO3 is heated to 500–600 °C at a heating rate of 5 °C/min in a protective atmosphere. When the preset temperature is reached, the argon gas is turned off, and hydrogen gas is introduced. After holding for a period of time (Mo9O26 is reduced for 15–45 min, Mo4O11 is reduced for 45–75 min), the MoO3 is slightly reduced by hydrogen to obtain substoichiometric molybdenum oxide. Substoichiometric molybdenum oxide is used for the characterization of Section 2.2 and the mixing and sintering of Section 2.3.

2.2. Thermal Stability Tests

The prepared Mo9O26 and Mo4O11 powders are subjected to synchronous differential scanning calorimetry (DSC) analysis and thermogravimetric analysis (TGA) on a synchronous thermal analyzer to explore the thermal stability of the oxides. In order to further explore the decomposition behavior and phase evolution of the two oxides at different temperatures, Mo9O26 and Mo4O11 are heat-treated in the temperature range of 700–1000 °C and 700–1100 °C for 1 h, respectively, using a tube furnace and corundum crucible (50 × 40 × 20 mm3). The isothermal stability test is carried out at an argon flow rate of 70 mL/min. The tubular furnace maintains the same heating rate (10 °C/min) as the synchronous thermal analysis. When the oxide melts at high temperatures and cools to room temperature to become bulky, the bulk is ground into powder using a mortar. Pure phase Mo9O26 and MoO3 (phase pure from XRD analysis; pure Mo9O26 is prepared by optimizing the hydrogen reduction process of Section 2.1, and MoO3 is the same as that used in Section 2.1) are used for heat treatment at 800–900 °C and 900 °C, respectively (conditions consistent with the isothermal thermal stability tests), as supplementary experiments to the isothermal thermal stability test experiments.

2.3. Sintered MoOx Ceramics

The substoichiometric molybdenum oxide powders prepared in Section 2.1 are mixed for sintering the ceramic bulks. By adjusting the ratio of substoichiometric molybdenum oxide, the O/Mo atomic ratio of the ceramic bulks can be controlled. In order to distinguish it from the unmixed powder, the mixed powder and ceramic bulk are expressed in the form of MoOx. For molybdenum oxide with an O/Mo atomic ratio of 2.9, there is no need to mix other substoichiometric molybdenum oxide to adjust the atomic ratio, so Mo9O26 powder is directly used as the raw material for sintering MoO2.9. Using a v-type mixer, Mo9O26 (62.3 wt.%), Mo4O11 (28.3 wt.%) and MoO2 (9.4 wt.%) are uniformly mixed to obtain MoO2.8 powder. Mo4O11 (80 wt.%, 70 wt.%) and MoO2 (20 wt.%, 30 wt.%) are uniformly mixed to obtain MoOx powder with O/Mo of 2.6 and 2.5. The speed of the v-type mixer is 50 rad/min, and the mixing time is 6 h. Then, the oxide is placed into a graphite mold with an inner diameter of 20 mm. The sintering process is carried out in the LABOX-675F SPS furnace (SINTER LAND INC., Niigata, Japan) at a pressure of 60 MPa for 15 min. The sintering temperature is between 700 and 800 °C, the heating rate is 100 °C/min, and the temperature in the furnace is detected using infrared temperature measuring equipment.

2.4. Characterization

The phase evolution and composition after heat treatment were examined by an X-ray diffractometer (XRD, Empyrean Alpha 1, Malvern Panalytical, Alemlo, The Netherlands) equipped with a diffracted beam monochromator using a Cu Kα radiation source. The micromorphology of the powders was analyzed by scanning electron microscope (SEM, Quanta 200 FEG, FEI, Hillsboro, OR, USA). The powders were sprayed with gold for better electronic conductivity. The TG–DSC measurements of the powders were carried out on a synchronous thermal analyzer (STA 449 F3, NETZSCH, Selbu, Germany) with a heating rate of 10 °C/min in argon (gas flow rate 100 mL/min) using alumina crucibles. The reflectivity was tested with a UV–VIS–NIR spectrophotometer (UV-3600, Shimadzu, Kyoto, Japan). The resistivity was measured with a four-point probe (RTS-8, Guangzhou Four Probe Technology Co., Ltd., Guangzhou, China).

3. Results and Discussion

3.1. Phase Evolution of Substoichiometric Molybdenum Oxide during Heat Treatment

The XRD patterns of the Mo9O26 and Mo4O11 powders prepared by reducing molybdenum trioxide are shown in Figure 1. Figure 1a shows that the powder is mainly Mo9O26 (89.4 wt.%) and contains a small amount of MoO2 (2.1 wt.%) and MoO3 (8.5 wt.%). Figure 1b shows that the powder is dominated by Mo4O11 (89.7 wt.%), which also contains a small amount of MoO2 (10.3 wt.%). The Mo9O26 and Mo4O11 powders mentioned below refer to these mixed powders.
Simultaneous TG–DSC curves for the Mo9O26 and Mo4O11 powders are shown in Figure 2. Figure 2a shows that the DSC curve contains three endothermic peaks at 774.1 °C, 811.3 °C and 1012.6 °C. There are three exothermic peaks at 805.1 °C, 885.6 °C and 930.1 °C, indicating that new phase crystallization may occur at these three temperatures. According to the TG curve, the initial mass reduction occurred at 766.1 °C. At approximately 1000 °C, the TG curve has an obvious anomaly, which corresponds to the endothermic peak at 1012.6 °C (Figure 2a). The final residual mass detected at 1200 °C is 27.7%. The Figure 2b DSC curve contains two endothermic peaks at 810.7 °C and 1037.9 °C. Exothermic peaks appear at 845 °C and 927.9 °C. According to the TG curve, the initial mass reduction occurred at 793.0 °C. At approximately 1030 °C, the TG curve, similar to Mo9O26, also has an obvious anomaly. This corresponds to the endothermic peak at 1037.9 °C (Figure 2b), which may be due to the violent sublimation [30]. The final residual mass detected at 1200 °C is 40.8%. According to the area method shown in Figure 2, the reaction enthalpy during the heat treatment is indicated.
Mo9O26 and Mo4O11 were subjected to isothermal heat treatment with reference to the temperature points obtained by the synchronous thermal analysis. As shown in Figure 3, the mass residual curve of substoichiometric molybdenum oxide after heat treatment is consistent with the TG curve. This same trend shows that the thermal decomposition behavior in the tubular furnace is consistent with that in the synchronous thermal analysis. Notably, as the heat treatment temperature increases, the powders become bulky when cooled to room temperature. The Mo9O26 powders form a bulk at 800 °C, and the Mo4O11 powders form a bulk at 840 °C (Figure 3). This is consistent with the judgment in the literature [24] that, above 800 °C, the Mo-O system will be in a liquid state.
Figure 4 shows the XRD spectrums of Mo9O26 and Mo4O11 after heat treatment at 750–950 °C and 800–1000 °C, respectively. Table 1 and Table 2 list the phase composition based on Figure 4. It can be observed from Table 1 that Mo9O26 still exists stably at 790 °C. MoO3 and MoO2 that originally existed in Mo9O26 disappeared, and Mo4O11 phase appeared, which may be the reaction (1):
3 MoO 3 + MoO 2   = Mo 4 O 11
This situation is similar to the method of Jie Dang et al. [31] to prepare Mo4O11. The slight decrease in mass should be attributed to the sublimation of MoO3. At 800 °C, MoO3 and Mo4O11 appear, and the content of Mo9O26 decreases significantly, indicating that Mo9O26 begins to fusion, and its enthalpy of fusion is 0.93 J/g. Between 820–840 °C, Mo9O26 gradually disappears, as shown in Figure 4a2. At 860 °C, Mo9O26 completely disappeared, and only Mo4O11, MoO3 and MoO2 three phases existed. New phase MoO2 is generated, which corresponds to the exothermic peak at 885.6 °C. At 900–950 °C, as MoO2 gradually becomes the main phase, the content of Mo4O11 decreases, and MoO3 has been completely sublimated in this temperature range.
As shown in Table 2, Mo4O11 is still stable at 830 °C. After heat treatment at 840 °C, the content of Mo4O11 decreases, and a trace of MoO3 appears, indicating that Mo4O11 begins to fuse at this temperature, and the enthalpy of fusion is 1.26 J/g. With the increasing temperature, Mo4O11 continues to decompose, MoO3 sustains to sublimate, and the content of MoO2 gradually increases. Finally, only MoO2 exists at 1000 °C. It is worth noting that there are thermodynamically unstable β-MoO3 (PDF#47-1320, not marked in Figure 4) and thermodynamically stable α-MoO3 (PDF#05-0508) during the decomposition process of substoichiometric molybdenum oxide. There is a trend of β-MoO3 to α-MoO3 with the increase in temperature. This transition from the thermodynamically unstable phase to the thermodynamically stable phase is accompanied by heat release, which is consistent with the exothermic peak present at approximately 930 °C in Figure 2. From the above results, we can reasonably obtain the thermal stability of Mo9O26 and Mo4O11. The former exists stably from room temperature to 790 °C, and the latter exists stably from room temperature to 830 °C.

3.2. Decomposition Process of Substoichiometric Molybdenum Oxide during Heat Treatment

In order to exclude the interference of Mo4O11 generated by Equation (1) during the heat treatment, high-purity Mo9O26 was heat treated at 800 °C and 900 °C. As can be observed from Figure 5a, Mo4O11 appeared after heat treatment. As shown in Figure 5b, MoO3 forms Mo4O11 when heat treated at 900 °C, that is, oxygen escape occurs in molybdenum oxide during heat treatment. The unassigned peaks in Figure 5b are Al2(MoO4)3 and Al2O3 impurities introduced by the alumina crucible.
According to the phase evolution (Figure 4) and the decrease in the O/Mo atomic ratio (Figure 5b), the decomposition process of Mo9O26 and Mo4O11 can be summarized as the following reactions:
Mo 9 O 26   = Mo 4 O 11 + 5 MoO 3
2 Mo 9 O 26   = 4 Mo 4 O 11 + 2 MoO 3   + O 2
Mo 4 O 11   = MoO 2 + 3 MoO 3
2 Mo 4 O 11   = 4 MoO 2 + 4 MoO 3   + O 2
Thermodynamic analysis software HSC6.0 was used to calculate the thermodynamics of the equations. Figure 6 shows the Gibbs free energies as a function of temperature for the four decomposition reactions possible for the two substoichiometric molybdenum oxides in the standard state (101.325 kPa), and it can be observed that all four reactions reach the thermodynamic condition at high temperature.
According to the research results in Section 3.1, only MoO2 exists above 1000 °C, the TG curve has significant jitter, and the DSC curve has an endothermic peak; it is reasonable to speculate that Mo4O11 will sublime rapidly above 1000 °C (Figure 2a,b sublimation 10.1% and 8.9%, respectively). For Mo4O11, it decomposes below 1000 °C according to Equation (4) or (5) (excluding the sublimated part above 1000 °C), and the final theoretical mass residue is 28.7% or 47.4%, respectively. The actual mass residue is 40.8% (Figure 2b), which is between Equations (4) and (5). This indicates that, during the Mo4O11 decomposition, both Equations (4) and (5) occur. According to the law of conservation of mass, it can be deduced that 64.4 wt.% of Mo4O11 decomposes according to Equation (5), and another 35.6 wt.% decomposes according to Equation (4). For Mo9O26, according to the Equation (2) or (3), the theoretical mass residues are 16.4% or 29.7%, respectively. Therefore, during the decomposition of Mo9O26, Equations (2) and (3) occur simultaneously with the former accounting for 15.1 wt.%, and the latter accounting for 84.9 wt.%. In summary, most of the mass loss of the two substoichiometric molybdenum oxides come from MoO3, and a few are due to the sublimation of Mo4O11 above 1000 °C.
Figure 7 and Figure 8 show the morphology of the Mo9O26 and Mo4O11 powders after heat treatment at 790–950 °C and 820–1100 °C, respectively. From Figure 7a,b, it can be found that the layered features of Mo9O26 disappear after heat treatment. As shown in Figure 7c,d, Mo9O26 begins to melt between 800–820 °C. The melting phenomenon can be observed in Figure 7d. There are obvious traces of droplets on the powders, and there is no particle adhesion on the surface of the droplets. The droplets are the residues of liquid MoO3 that are not volatilized completely. As the temperature continues to increase to 860 °C (Figure 7e), subgrains are found on the powder surface, and the powder again shows distinct layered features (Figure 7d,e). The presence of layered features in Figure 7d,e may be due to the Mo9O26 of the layered structure [32] with high energy at the interlaminar interface, such that the decomposition reaction occurs here first, and the powder is turned into thinner particles. This transformation is very similar to the crackling core model (CCM). During the temperature increase from 860 °C (Figure 7e) to 950 °C (Figure 7f), the subgrains on the surface of the powder begin to grow and increase significantly, and the plate-like particles are completely covered at 950 °C. From the perspective of phase transition between 860–950 °C, MoO2 continues to increase, and Mo4O11 gradually decreases, so it can be judged that the small particles added on the surface are newly formed MoO2 phases. This indicates that the transition from Mo4O11 to MoO2 during heat treatment is an external-to-internal reaction.
As shown in Figure 8a,b, Mo4O11 grows from an irregular polygonal structure to a regular geometry in which small disk-shaped particles are embedded. This adhesion of different particles may be caused by the low melting point eutectic [33]. At 830 °C (Figure 8c), the Mo4O11 powder is nearly spherical. Decomposition begins at 840 °C (Figure 8d), and small particles begin to appear on the surface. At 900 °C (Figure 8e), small particles increase and grow significantly and coat on the powder surfaces. From the phase change, it can be observed that, between 840–900 °C, the MoO2 content continues to increase, so these gradually increasing small particles are MoO2. At 1100 °C (Figure 8f), the powder is completely transformed into polygonal MoO2 (the XRD pattern of heat treatment at 1100 °C is consistent with that at 1000 °C, so it is not shown in Figure 4). It can be found that the two kinds of Mo4O11 are (1) obtained by thermal decomposition of Mo9O26 at 860 °C and above, and (2) obtained by hydrogen reduction and have the same transformation behavior, which fits the chemical vapour transport (CVT) model from outside to inside; the former is shown in Figure 7e,f, and the latter is shown in Figure 8e,f. The transformation of molybdenum oxide undergoes CCM and CVT, which is different from the view of Werner V. Schulmeyer et al. [34], who think that MoO2 is obtained by hydrogen reduction only through the CVT model, which may be because they have not found that Mo9O26 or the conditions of the two reactions (heat treatment or hydrogen reduction) are different.
The stability and decomposition mechanism of substoichiometric molybdenum oxide are shown in Figure 9. Below 800 °C, Mo9O26 can exist stably. Then, Mo9O26 cracks through the CCM at 800 °C and begins to convert to Mo4O11. Until 840 °C, Mo9O26 almost decomposes completely. Mo4O11 is stable below 840 °C and forms MoO2 subgrains on the surface by CVT at 840 °C; then, the subgrains increase and grow, and finally replace Mo4O11 completely from the surface to the interior.
Comparing the experimental results with the Mo-O phase diagram (Figure 10), we find that the decomposition process of Mo4O11 is consistent with the phase diagram, transforming into MoO2 and MoO3 at high temperatures. The decomposition of Mo9O26 is not consistent with the phase diagram; it is directly decomposed into Mo4O11 and MoO3, and Mo8O23 does not appear during the decomposition process. The decomposition temperatures of Mo4O11 and Mo9O26 phases in the phase diagram are lower than those of the isothermal thermal stability experiments, but they are consistent with the temperature at which the first endothermic peak appears in the DSC curve (Figure 2).

3.3. Phase Composition, Optical and Electrical Properties of MoOx Ceramic Bulks

The Archimedes drainage method was used to determine the relative density of four different x ceramic materials prepared by SPS sintering, which were higher than 96.4%. The phase composition of the ceramic bulk is shown in Figure 11. The phase of MoO2.9 (Sample A) and MoO2.8 (Sample B) did not change before and after sintering. After sintering with MoO2.6 (Sample C) and MoO2.5 (Sample D), Mo4O11 increased by 10 wt.%, and MoO2 decreased by 10 wt.%. C and D rose to MoO2.7 and MoO2.6, respectively, which should be due to the increase in MoO2 content that promoted its reaction with the undetected high x (O/Mo atomic ratio, x > 2.75) molybdenum oxide in the raw material, thus producing Mo4O11.
The optical and electrical properties of the MoOx bulks are characterized by reflectivity and resistivity tests. The reflectivity and resistivity are shown in Figure 12. At the most sensitive area of the human eye (550 nm), the reflectivity is between 6.3 to 7.8%, among which the reflectivity of MoO2.8 is the lowest at 6.3%. With the decrease in x, the resistivity decreases from 10−1 to 10−3 Ωcm, and the measured value is close to the value recorded in the literature [36]. From the change in resistivity, MoO2 with low resistivity significantly affects the conductivity of the MoOx ceramics. MoO2.9 contains only Mo9O26, which exhibits the highest resistivity. The resistivity of the other three MoOx ceramics decreases significantly with the addition of MoO2. Through the analysis of optical and electrical properties, it can be determined that MoOx material is a low reflective conductive oxide material with great application prospects.

4. Conclusions

The thermal stability and decomposition mechanism of Mo9O26 and Mo4O11 are systematically studied. Under an argon atmosphere, Mo9O26 has good thermal stability below 790 °C, gradually decomposes into Mo4O11 and MoO3 between 800–840 °C, and the transformation form is similar to the CCM. At 860–950 °C, Mo4O11 generated by thermal decomposition decomposes into MoO2 and MoO3. Mo4O11 has good thermal stability below 830 °C, is decomposed into MoO2 and MoO3 from 840 °C and completely transforms into MoO2 at 1000 °C. This transition is consistent with the CVT model. The decomposition of molybdenum oxide is the process of oxygen loss with the increase in temperature. In a word, during the heat treatment, the transformation of molybdenum oxide undergoes CCM and CVT successively. The relative density of ceramic bulks obtained by SPS sintering can reach 96.4%. All four MoOx have low reflectivity that ranges from 6.3–7.8%, especially MoO2.8 at 6.3%. The resistivity of MoOx decreases from 10−1 to 10−3 Ωcm with the decrease in x.

Author Contributions

Formal analysis, K.Y., Y.L., Q.L., Y.C. and C.C.; Investigation, Y.Q. and Y.C.; Resources, J.Z., B.S. and J.H.; Writing—original draft, Y.Q.; Writing—review & editing, K.Y. and Y.C.; Supervision, Q.L. and J.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Henan province science and technology research plan project (No. 212102210594), the National Key Research and Development Program of China (No. 2021YFB3600803), and the Zhengzhou Collaborative Innovation Major Project (125/23240001).

Data Availability Statement

The data presented in this study are available on request from the corresponding authors.

Acknowledgments

The authors would like to acknowledge the support received from the Henan province science and technology research plan project (No. 212102210594), the National Key Research and Development Program of China (No. 2021YFB3600803), and the Zhengzhou Collaborative Innovation Major Project (125/23240001).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yao, D.D.; Rani, R.A.; O’Mullane, A.P.; Kalantar-Zadeh, K.; Ou, J.Z. Enhanced Coloration Efficiency for Electrochromic Devices based on Anodized Nb2O5/Electrodeposited MoO3 Binary Systems. J. Phys. Chem. C 2014, 118, 10867–10873. [Google Scholar] [CrossRef]
  2. Hanlon, D.; Backes, C.; Higgins, T.M.; Hughes, M.; O’Neill, A.; King, P.; McEvoy, N.; Duesberg, G.S.; Mendoza Sanchez, B.; Pettersson, H.; et al. Production of molybdenum trioxide nanosheets by liquid exfoliation and their application in high-performance supercapacitors. Chem. Mater. 2014, 26, 1751–1763. [Google Scholar] [CrossRef]
  3. Xiang, D.; Han, C.; Zhang, J.; Chen, W. Gap states assisted MoO3 nanobelt photodetector with wide spectrum response. Sci. Rep. 2014, 4, 4891. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Yu, X.; Marks, T.J.; Facchetti, A. Metal oxides for optoelectronic applications. Nat. Mater. 2016, 15, 383–396. [Google Scholar] [CrossRef] [PubMed]
  5. Greiner, M.T.; Chai, L.; Helander, M.G.; Tang, W.-M.; Lu, Z.-H. Metal/metal-oxide interfaces: How metal contacts affect the work function and band structure of MoO3. Adv. Funct. Mater. 2013, 23, 215–226. [Google Scholar] [CrossRef]
  6. Liu, J.; Shao, S.; Fang, G.; Meng, B.; Xie, Z.; Wang, L. High-efficiency inverted polymer solar cells with transparent and work-function tunable MoO3-Al composite film as cathode buffer layer. Adv. Mater. 2012, 24, 2774–2779. [Google Scholar] [CrossRef]
  7. Mai, L.Q.; Hu, B.; Chen, W.; Qi, Y.Y.; Lao, C.S.; Yang, R.S.; Dai, Y.; Wang, Z.L. Lithiated MoO3 nanobelts with greatly improved performance for lithium batteries. Adv. Mater. 2007, 19, 3712–3716. [Google Scholar] [CrossRef]
  8. Choi, S.H.; Kang, Y.C. Crumpled graphene-molybdenum oxide composite powders: Preparation and application in lithium-ion batteries. ChemSusChem 2014, 7, 523–528. [Google Scholar] [CrossRef]
  9. Noerochim, L.; Wang, J.-Z.; Wexler, D.; Chao, Z.; Liu, H.-K. Rapid synthesis of free-standing MoO3/graphene films by the microwave hydrothermal method as cathode for bendable lithium batteries. J. Power Source 2013, 228, 198–205. [Google Scholar] [CrossRef] [Green Version]
  10. Wang, S.; Guo, H.; Song, Y.; Ma, G.; Li, S.; Zhang, L.; Shao, X. Lower reflective TFT materials and technology innovation. SID Symp. Dig. Tech. Pap. 2017, 48, 1478–1481. [Google Scholar] [CrossRef]
  11. Schmidt, H.; Koestenbauer, H.; Koestenbauer, J.; Linke, C.; Franzke, E.; Winkler, J. Sputtered molybdenum-oxide for anti-reflection layers in displays: Optical properties and thermal stability. SID Symp. Dig. Tech. Pap. 2018, 49, 225–229. [Google Scholar] [CrossRef]
  12. Kumar, M.; Cho, E.-C.; Prodanov, M.F.; Kang, C.; Srivastava, A.K.; Yi, J. MoOx work function, interface structure, and thermal stability analysis of ITO/MoOx/a-Si(i) stacks for hole-selective silicon heterojunction solar cells. Appl. Surf. Sci. 2021, 553, 149552. [Google Scholar] [CrossRef]
  13. Battaglia, C.; Yin, X.; Zheng, M.; Sharp, I.D.; Chen, T.; McDonnell, S.; Azcatl, A.; Carraro, C.; Ma, B.; Maboudian, R.; et al. Hole selective MoOx contact for silicon solar cells. Nano Lett. 2014, 14, 967–971. [Google Scholar] [CrossRef]
  14. Qiu, W.; Müller, R.; Voroshaz, E.; Conings, B.; Carleer, R.; Boyen, H.-G.; Turbiez, M.; Froyen, L.; Heremans, P.; Hadipour, A. Nafion-modified MoOx as effective room-temperature hole injection layer for stable, high-performance inverted organic solar cells. ACS Appl. Mater. Interfaces 2015, 7, 3581–3589. [Google Scholar] [CrossRef] [PubMed]
  15. Wu, W.; Bao, J.; Liu, Z.; Lin, W.; Yu, X.; Cai, L.; Liu, B.; Song, J.; Shen, H. Multilayer MoOx/Ag/MoOx emitters in dopant-free silicon solar cells. Mater. Lett. 2017, 189, 86–88. [Google Scholar] [CrossRef]
  16. Li, Y.; Wang, D.; An, Q.; Ren, B.; Rong, Y.; Yao, Y. Flexible electrode for long-life rechargeable sodium-ion batteries: Effect of oxygen vacancy in MoO3−x. J. Mater. Chem. A 2016, 4, 5402–5405. [Google Scholar] [CrossRef]
  17. Meduri, P.; Pendyala, C.; Kumar, V.; Sumanasekera, G.U.; Sunkara, M.K. Hybrid Tin oxide nanowires as stable and high capacity anodes for Li-ion batteries. Nano Lett. 2009, 9, 612–616. [Google Scholar] [CrossRef]
  18. Scanlon, D.O.; Watson, G.W.; Payne, D.J.; Atkinson, G.R.; Egdell, R.G.; Law, D.S.L. Theoretical and experimental study of the electronic structures of MoO3 and MoO2. J. Phys. Chem. C 2010, 114, 4636–4645. [Google Scholar] [CrossRef]
  19. Luo, Z.; Miao, R.; Huan, T.D.; Mosa, I.M.; Poyraz, A.S.; Zhong, W.; Cloud, J.E.; Kriz, D.A.; Thanneeru, S.; He, J.; et al. Mesoporous MoO3–x material as an efficient electrocatalyst for hydrogen evolution reactions. Adv. Energy Mater. 2016, 6, 1–11. [Google Scholar] [CrossRef]
  20. Wang, J.; Wang, Z.; Huang, B.; Ma, Y.; Liu, Y.; Qin, X.; Zhang, X.; Dai, Y. Oxygen vacancy induced band-gap narrowing and enhanced visible light photocatalytic activity of ZnO. ACS Appl. Mater. Interfaces 2012, 4, 4024–4030. [Google Scholar] [CrossRef]
  21. Song, G.; Hao, J.; Liang, C.; Liu, T.; Gao, M.; Cheng, L.; Hu, J.; Liu, Z. Degradable molybdenum oxide nanosheets with rapid clearance and efficient tumor homing capabilities as a therapeutic nanoplatform. Angew. Chem. Int. Ed. 2016, 55, 2122–2126. [Google Scholar] [CrossRef] [PubMed]
  22. Balendhran, S.; Deng, J.; Ou, J.Z.; Walia, S.; Scott, J.; Tang, J.; Wang, K.L.; Field, M.R.; Russo, S.; Zhuiykov, S.; et al. Enhanced charge carrier mobility in two-dimensional high dielectric molybdenum oxide. Adv. Mater. 2013, 25, 109–114. [Google Scholar] [CrossRef] [PubMed]
  23. Alsaif, M.M.Y.A.; Chrimes, A.F.; Daeneke, T.; Balendhran, S.; Bellisario, D.O.; Son, Y.; Field, M.R.; Zhang, W.; Nili, H.; Nguyen, E.P.; et al. High-performance field effect transistors using electronic inks of 2D molybdenum oxide nanoflakes. Adv. Funct. Mater. 2016, 26, 91–100. [Google Scholar] [CrossRef]
  24. Chang, L.L.Y.; Phillips, B. Phase relations in refractory metal-oxygen systems. J. Am. Ceram. Soc. 1969, 52, 527–533. [Google Scholar] [CrossRef]
  25. Cho, J.; Nawal, N.; Hadipour, A.; Payo, M.R.; van der Heide, A.; Radhakrishnan, H.S.; Debucquoy, M.; Gordon, I.; Szlufcik, J.; Poortmans, J. Interface analysis and intrinsic thermal stability of MoOx based hole-selective contacts for silicon heterojunction solar cells. Sol. Energy Mater Sol. Cells 2019, 201, 110074. [Google Scholar] [CrossRef]
  26. Cao, S.; Li, J.; Lin, Y.; Pan, T.; Du, G.; Zhang, J.; Yang, L.; Chen, X.; Lu, L.; Min, N.; et al. Interfacial behavior and stability analysis of p-type crystalline silicon solar cells based on hole-selective MoOx/metal contacts. Sol. RRL 2019, 3, 1900274. [Google Scholar] [CrossRef]
  27. Mallem, K.; Kim, Y.J.; Hussain, S.Q.; Dutta, S.; Le, A.H.T.; Ju, M.; Park, J.; Cho, Y.H.; Kim, Y.; Cho, E.-C.; et al. Molybdenum oxide: A superior hole extraction layer for replacing p-type hydrogenated amorphous silicon with high efficiency heterojunction Si solar cells. Mater. Res. Bull. 2019, 110, 90–96. [Google Scholar] [CrossRef]
  28. Greiner, M.T.; Chai, L.; Helander, M.G.; Tang, W.-M.; Lu, Z.-H. Transition metal oxide work functions: The influence of cation oxidation state and oxygen vacancies. Adv. Funct. Mater. 2012, 22, 4557–4568. [Google Scholar] [CrossRef]
  29. Gregory, G.; Wilson, M.; Ali, H.; Davis, K.O. Thermally stable molybdenum oxide hole-selective contacts deposited using spatial atomic layer deposition. In Proceedings of the 2018 IEEE 7th World Conference on Photovoltaic Energy Conversion (WCPEC) (A Joint Conference of 45th IEEE PVSC), Waikoloa Village, HI, USA, 10–15 June 2018. [Google Scholar]
  30. Zhang, F. Interpretation of causing unusual thermogravimetric curve and its improvement method. Exp. Technol. Manag. 2016, 33, 32–37. [Google Scholar] [CrossRef]
  31. Dang, J.; Zhang, G.-H.; Wang, L.; Chou, K.-C. Study on hydrogen reduction of Mo4O11. Int. J. Refract. Met. Hard Mater. 2015, 51, 275–281. [Google Scholar] [CrossRef]
  32. Magnéli, A. The crystal structures of Mo9O26 (β′-molybdenum oxide) and Mo8O23 (β-molybdenum oxide). Acta Chem. Scand. 1948, 2, 501–517. [Google Scholar] [CrossRef]
  33. Wang, L.; Xue, Z.L.; Huang, A.; Wang, F.F. Mechanism and kinetic study of reducing MoO3 to MoO2 with CO-15 vol % CO2 mixed gases. ACS Omega 2019, 4, 20036–20047. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Schulmeyer, W.V.; Ortner, H.M. Mechanisms of the hydrogen reduction of molybdenum oxides. Int. J. Refract. Met. Hard Mater. 2002, 20, 261–269. [Google Scholar] [CrossRef]
  35. Zhang, C.; Gao, M.C.; Yang, Y.; Zhang, F. Thermodynamic modeling and first-principles calculations of the Mo–O system. Calphad 2014, 45, 178–187. [Google Scholar] [CrossRef]
  36. Gruber, H.; Krautz, E. Untersuchungen der elektrischen Leitfähigkeit und des Magnetowiderstandes im System Molybdän-Sauerstoff. Phys. Status Solidi (A) Appl. Res. 1980, 62, 615–624. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of Mo9O26 (a) and Mo4O11 (b).
Figure 1. XRD patterns of Mo9O26 (a) and Mo4O11 (b).
Materials 16 02841 g001
Figure 2. TG–DSC curves recorded during heating of the Mo9O26 (a) and Mo4O11 (b) powders up to 1200 °C at a rate of 10 °C/min in flowing argon.
Figure 2. TG–DSC curves recorded during heating of the Mo9O26 (a) and Mo4O11 (b) powders up to 1200 °C at a rate of 10 °C/min in flowing argon.
Materials 16 02841 g002
Figure 3. Residual mass of substoichiometric molybdenum oxides after heat treatment.
Figure 3. Residual mass of substoichiometric molybdenum oxides after heat treatment.
Materials 16 02841 g003
Figure 4. XRD patterns of the Mo9O26 (a1,a2) and Mo4O11 (b1,b2) powders after heat treatment.
Figure 4. XRD patterns of the Mo9O26 (a1,a2) and Mo4O11 (b1,b2) powders after heat treatment.
Materials 16 02841 g004
Figure 5. XRD patterns of the Mo9O26 (a) and MoO3 (b) powders after heat treatment.
Figure 5. XRD patterns of the Mo9O26 (a) and MoO3 (b) powders after heat treatment.
Materials 16 02841 g005
Figure 6. Changes in Gibbs free energy of thermal decomposition of Mo9O26 and Mo4O11 with temperature under standard state.
Figure 6. Changes in Gibbs free energy of thermal decomposition of Mo9O26 and Mo4O11 with temperature under standard state.
Materials 16 02841 g006
Figure 7. Surface morphologies of the Mo9O26 raw powders (a) and them after heat treated at 790 °C (b), 800 °C (c), 820 °C (d), 860 °C (e) and 950 °C (f) for 1 h in flowing argon.
Figure 7. Surface morphologies of the Mo9O26 raw powders (a) and them after heat treated at 790 °C (b), 800 °C (c), 820 °C (d), 860 °C (e) and 950 °C (f) for 1 h in flowing argon.
Materials 16 02841 g007
Figure 8. Surface morphologies of the Mo4O11 raw powders (a) and them after heat treated at 820 °C (b), 830 °C (c), 840 °C (d), 900 °C (e) and 1100 °C (f) for 1 h in flowing argon.
Figure 8. Surface morphologies of the Mo4O11 raw powders (a) and them after heat treated at 820 °C (b), 830 °C (c), 840 °C (d), 900 °C (e) and 1100 °C (f) for 1 h in flowing argon.
Materials 16 02841 g008
Figure 9. Thermal stability and decomposition mechanism of substoichiometric molybdenum oxides.
Figure 9. Thermal stability and decomposition mechanism of substoichiometric molybdenum oxides.
Materials 16 02841 g009
Figure 10. Mo–O binary phase diagram [35].
Figure 10. Mo–O binary phase diagram [35].
Materials 16 02841 g010
Figure 11. XRD patterns of A (a), B (b), C (c) and D (d) before and after sintering.
Figure 11. XRD patterns of A (a), B (b), C (c) and D (d) before and after sintering.
Materials 16 02841 g011
Figure 12. Reflectivity (a) and resistivity (b) of the MoOx ceramics. The resistivity also includes the reference values of metal Mo and MoO2 in the literature [36].
Figure 12. Reflectivity (a) and resistivity (b) of the MoOx ceramics. The resistivity also includes the reference values of metal Mo and MoO2 in the literature [36].
Materials 16 02841 g012
Table 1. Phase compositions of the Mo9O26 powders after heat treatment.
Table 1. Phase compositions of the Mo9O26 powders after heat treatment.
Heat Treatment Temperature (°C)Phase Composition
750Mo9O26(main), Mo4O11
780Mo9O26(main), Mo4O11
790Mo9O26(main), Mo4O11
800Mo9O26, Mo4O11, MoO3
820Mo4O11, MoO3, Mo9O26
840Mo4O11, MoO3, Mo9O26(trace)
860Mo4O11, MoO3, MoO2
900Mo4O11(main), MoO3, MoO2
950MoO2(main), Mo4O11
Table 2. Phase compositions of the Mo4O11 powders after heat treatment.
Table 2. Phase compositions of the Mo4O11 powders after heat treatment.
Heat Treatment Temperature (°C)Phase Composition
800Mo4O11(main), MoO2
820Mo4O11(main), MoO2
830Mo4O11(main), MoO2
840Mo4O11, MoO2, MoO3(trace)
850Mo4O11, MoO2, MoO3(trace)
860Mo4O11, MoO2
900MoO2(main), Mo4O11
990MoO2, Mo4O11(trace)
1000MoO2
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

Qing, Y.; Yang, K.; Chen, Y.; Zhu, J.; Li, Y.; Chen, C.; Li, Q.; Sun, B.; He, J. Thermal Stability, Optical and Electrical Properties of Substoichiometric Molybdenum Oxide. Materials 2023, 16, 2841. https://doi.org/10.3390/ma16072841

AMA Style

Qing Y, Yang K, Chen Y, Zhu J, Li Y, Chen C, Li Q, Sun B, He J. Thermal Stability, Optical and Electrical Properties of Substoichiometric Molybdenum Oxide. Materials. 2023; 16(7):2841. https://doi.org/10.3390/ma16072841

Chicago/Turabian Style

Qing, Yubin, Kaijun Yang, Yaofeng Chen, Jinpeng Zhu, Yujing Li, Chong Chen, Qingkui Li, Benshuang Sun, and Jilin He. 2023. "Thermal Stability, Optical and Electrical Properties of Substoichiometric Molybdenum Oxide" Materials 16, no. 7: 2841. https://doi.org/10.3390/ma16072841

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