Next Article in Journal
Catalytic Behaviors of Supported Cu, Ni, and Co Phosphide Catalysts for Deoxygenation of Oleic Acid
Previous Article in Journal
Heterogeneous Synergistic Catalysis for Promoting Aza-Michael–Henry Tandem Reaction for the Synthesis of Chiral 3-Nitro-1,2-Dihydroquinoline
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Cu0.4Co0.6MoO4 Nanorods Supported on Graphitic Carbon Nitride as a Highly Active Catalyst for the Hydrolytic Dehydrogenation of Ammonia Borane

1
School of Chemical Engineering and Light Industry, Guangdong University of Technology, Guangzhou 510006, China
2
School of Chemistry and Materials Engineering, Huizhou University, Huizhou 516007, China
*
Authors to whom correspondence should be addressed.
Catalysts 2019, 9(9), 714; https://doi.org/10.3390/catal9090714
Submission received: 22 July 2019 / Revised: 18 August 2019 / Accepted: 21 August 2019 / Published: 24 August 2019
(This article belongs to the Section Catalytic Materials)

Abstract

:
As a typical chemical hydride, ammonia borane (AB) has received extensive attention because of its safety and high hydrogen storage capacity. The aim of this work was to develop a cost-efficient and highly reactive catalyst for hydrolyzing AB. Herein, we synthesized a series of CuxCo1–xMoO4 dispersed on graphitic carbon nitride (g-C3N4) to dehydrogenate AB. Among those CuxCo1–xMoO4/g-C3N4 catalysts, Cu0.4Co0.6MoO4/g-C3N4 exhibited the highest site time yield (STY) value of 75.7 m o l H 2 m o l c a t 1 m i n 1 with a low activation energy of 14.46 kJ mol−1. The STY value for Cu0.4Co0.6MoO4/g-C3N4 was about 4.3 times as high as that for the unsupported Cu0.4Co0.6MoO4, indicating that the g-C3N4 support plays a crucial role in improving the catalytic activity. Considering its low cost and high catalytic activity, our Cu0.4Co0.6MoO4/g-C3N4 catalyst is a strong candidate for AB hydrolysis for hydrogen production in practical applications.

1. Introduction

With the rapid development of social economy and technology, the consumption of fossil fuels keeps growing, leading to the destruction of the environment and ecology [1,2]. Hydrogen energy, generating H2O after reaction, can effectively solve the contradiction in the current energy structure between developing need and the environment [3]. To realize the vision of hydrogen power popularization, problems of hydrogen storage and acquisition must be overcome [4]. Chemical hydrogen storage, one of the most popular approaches, is the process by which hydrogen is trapped in a liquid or solid and can be extracted in due course. Ammonia borane (AB), with its high hydrogen density (19.6%), low molar molecular weight (30.7 g mol−1), and high stability, is a high-profile hydrogen storage material [5,6,7,8]. When appropriate catalysts are used, the activation energy of the reaction is reduced, and the AB produces a large amount of hydrogen under mild conditions according to the following reaction [9,10]:
NH3BH3 + 2H2O → NH4BO2 + 3H2
In the literature, great efforts have been made to develop noble-metal-based catalysts [11]. However, their large-scale applications are remarkably limited by their high costs. Although the costs of bimetallic catalysts composed of both noble and nonnoble metals, such as Au–Ni, Ru–Ni, etc., are reduced in contrast to those of the noble metal catalysts, they are still high [12,13]. Thus, it is important for us to further reduce the cost of catalysts by developing noble-metal-free catalysts. On the other hand, the introduction of supports is one of the most efficient strategies for restraining the agglomeration of nanoparticles during a catalytic reaction. Due to its unique structure, inexpensiveness, and high specific surface area, carbon nitride (g-C3N4) has proved to be an ideal support in the field of heterogeneous catalysis, such as in photocatalysis, electrocatalysis, and the oxidation of toluene [14,15,16]. In recent years, there have been several successful examples of the application of g-C3N4 as a catalyst support in AB dehydrogenation. For example, Guo et al. designed g-C3N4-supported Au–Co nanoparticles with a site time yield (STY) value of 28.4 m o l H 2 m o l c a t 1 m i n 1 [17]. In contrast, the STY value of Au–Co nanoparticles without g-C3N4 support was only ca. 14 m o l H 2 m o l c a t 1 m i n 1 . Fan et al. synthesized Rh nanoparticles with an average size of 3.1 nm which dispersed onto g-C3N4. The STY value for the resultant catalysts was 969 m o l H 2 m o l c a t 1 m i n 1 [18]. Hamza et al. reported that the STY value of AgPd/g-C3N4 was 94.1 m o l H 2 m o l c a t 1 m i n 1 [19]. Jia et al. synthesized Pb(0) nanoparticles anchored into g-C3N4 with chitosan, and their STY value was 27.7 m o l H 2 m o l c a t 1 m i n 1 at 30 °C [20]. Fan et al. synthesized AgCo bimetallic nanoparticles supported on g-C3N4, and the STY value of Ag0.1Co0.9/g-C3N4 was 249.02 m o l H 2 m o l c a t 1 m i n 1 [21]. It should be noted that the applications of g-C3N4 as a support in AB dehydrogenation are still rarely reported. In addition, the active components were noble metal or noble-metal-containing alloys in those reports. Therefore, it is important to explore some other types of active components supported on g-C3N4 and investigate their activity in AB dehydrogenation.
Herein, we report a series of CuxCo1–xMoO4 supported on g-C3N4 for ammonia borane hydrolysis. As far as we know, a noble-metal-free catalyst supported on g-C3N4 towards AB hydrolysis has not been reported previously. In this work, we found that the catalytic activity of the catalysts was closely related with the molar ratio of Co to Cu in CuxCo1–xMoO4/g-C3N4. It was discovered that Cu0.4Co0.6MoO4/g-C3N4 shows the best catalytic activity as its STY value can reach 75.7 molH2 molcat−1 min−1, which is better than the properties of most non-precious metals reported in the literature. In addition, Cu0.4Co0.6MoO4/g-C3N4 exhibited higher catalytic activity in contrast to the unsupported counterpart, Cu0.4Co0.6MoO4.

2. Results and Discussion

Figure 1 shows the Powder X-ray diffraction (XRD) patterns of g-C3N4 before loading and Cu0.4Co0.6MoO4/g-C3N4. Evidently, the peaks of g-C3N4, CuMoO4, and CoMoO4 can be found in Cu0.4Co0.6MoO4/g-C3N4. The diffraction peak at 27.7° can be observed in both XRD patterns and is the characteristic peak of g-C3N4. The diffraction peaks of Cu0.4Co0.6MoO4/g-C3N4 at 23.4°, 36.5°, 38.6°, 43.1°, and 52.9° correspond to the (110), (002), (200), ( 1 ¯ 1 ¯ 2), and (1 3 ¯ 0) planes of CuMoO4 (PDF#26-0546). The peaks at 26.5°, 40.2°, and 46.3° correspond to the (002), (003), and (403) planes of CoMoO4 (PDF#21-0868). No other peaks of impurities can be observed in the XRD pattern, indicating that the composites mainly contain C3N4, CoMoO4, and CuMoO4. For comparison, a series of CuxCo1–xMoO4/g-C3N4 were synthesized, and their XRD patterns are shown in Figure S1.
The Fourier transform infrared (FTIR) spectra of Cu0.4Co0.6MoO4, Cu0.4Co0.6MoO4/g-C3N4, and pure g-C3N4 are shown in Figure 2. For pure g-C3N4 powder, the main peaks at 815, 1240, 1316, 1400, 1458, 1573, and 1640 cm−1 can be seen, due to the stretching vibration modes of C=N and C–N, which is similar to previous reports [20,22,23]. For Cu0.4Co0.6MoO4/g-C3N4, the bands in the range of 1240–1640 cm−1 correspond to the bonds of pure g-C3N4. In contrast, the bands at 937, 615, and 475 cm−1 are assigned to the stretching vibrations of the t-MoO4 group, Mo=O stretching, and Mo–O–Mo bending vibrations [24,25,26]. This result further confirms that Cu0.4Co0.6MoO4 nanorods were successfully loaded onto g-C3N4. The FTIR spectra of different CuxCo1–xMoO4/g-C3N4 samples are shown in Figure S2 and are similar to that of Cu0.4Co0.6MoO4/g-C3N4.
The field-emission scanning electron microscope (FESEM) images and Energy Dispersive Spectrometer (EDS) spectra of the catalysts are shown in Figure 3 and Figure S3, respectively. The morphology of g-C3N4 in Figure 3a is similar to that in a previous report [6]. The unsupported Cu0.4Co0.6MoO4 is shown in Figure 3b,c, which shows aggregated nanorods with a diameter of ca. 500 nm. We note that most of the rods agglomerate together and form large aggregates, leading to a reduction in the contact area of the catalyst and substrate in the catalysis. To disperse Cu0.4Co0.6MoO4 well and increase the number of exposed active sites, we introduced g-C3N4 into solution during the synthesis process and successfully obtained the Cu0.4Co0.6MoO4/g-C3N4 material (Figure 3d,e). The rods could be easily found on the surface of g-C3N4. The diameter of the rods was still ca. 500 nm, and the length was ca. 5–10 μm. This observation indicates that the introduction of the g-C3N4 into the synthetic system did not change the morphology and size of the Cu0.4Co0.6MoO4 nanorods. However, their dispersion was remarkably improved. EDS analysis was carried out on the nanorods, and the results are shown in Figure 3f. The elements C, N, O, Mo, Co, and Cu were detected. Evidently, the elements C and N come from the g-C3N4 material, while the elements O, Mo, Co, and Cu result from the Cu0.4Co0.6MoO4. As displayed in Figure S3, the relative peak heights of Co to Cu decrease as x increases. The molar ratios of the different elements are shown in Table S1, indicating that the actual ratios of Co, Cu, and Mo in CuxCo1–xMoO4/g-C3N4 are close to the theoretical values.
The valence states of the elements in Cu0.4Co0.6MoO4/g-C3N4 were measured by X-ray photoelectron spectrometer (XPS), and the results are shown in Figure 4. The peaks at 284.4 eV and 287.7 eV correspond to C 1s, assigned to carbon atoms on the surface and coordination between carbon atoms and nitrogen atoms, respectively [23]. The N 1s spectrum separated into two peaks at 398.4 eV and 400.1 eV, which were assigned to sp2-hybridized nitrogen in triazine rings (C–N=C) and tertiary nitrogen (N–(C)3) groups [27]. The Co 2p spectrum shows two peaks at 780.9 eV and 796.1 eV of the Co 2p3/2 and Co 2p1/2 energy levels, demonstrating that Co is divalent [28]. At the same time, the peaks of Cu 2p at 931.8 eV and 951.7 eV imply the existence of Cu2+. Figure 4e is the XPS spectrum of Mo 3d. Two peaks at 232.2 eV and 235.3 eV imply that the element Mo is present as Mo6+ [29]. The peak at 531.3 eV in the O 1s spectrum illustrates the existence of O2− [30]. Considering the XRD, SEM, and XPS results together, we can conclude that our obtained product is g-C3N4-supported Cu0.4Co0.6MoO4.
The catalytic activity levels of g-C3N4, Cu0.4Co0.6MoO4, and Cu0.4Co0.6MoO4/g-C3N4 in the hydrolysis of ammonia borane were tested at 293 K. It can be seen in Figure 5 that g-C3N4 had no catalytic activity in the hydrolytic reaction. In addition, the catalytic activity of unsupported Cu0.4Co0.6MoO4 was significantly lower than that of Cu0.4Co0.6MoO4/g-C3N4. The STY values were 75.7 m o l H 2 m o l c a t 1 m i n 1 for Cu0.4Co0.6MoO4/g-C3N4 and 17.6 m o l H 2 m o l c a t 1 m i n 1 for the unsupported Cu0.4Co0.6MoO4. As discussed above, the introduction of g-C3N4 into the catalysts can help to disperse active materials and thereby improve the catalytic property tremendously. Notably, the molar ratio of hydrogen to ammonia borane reached 3, demonstrating that complete hydrogen release can be achieved when Cu0.4Co0.6MoO4/g-C3N4 acts as a catalyst. Figure 6 illustrates catalytic hydrolysis on the unsupported Cu0.4Co0.6MoO4 and Cu0.4Co0.6MoO4/g-C3N4 clearly. We also tried to synthesize ZnMoO4 by a similar strategy to measure the catalysis activity of other molybdates. The result that there was no catalysis activity in ZnMoO4 illustrates that molybdates have no reactivity. However, according to the literature, molybdates can serve as a Lewis acid and conduce the adsorption of OH on the catalyst surface, which would be of benefit to the hydrolysis reaction [31].
In consideration of variation of the molar ratio of cobalt to copper, we synthesized a series of CuxCo1–xMoO4/g-C3N4 (x = 0, 0.2, 0.4, 0.6, 0.8, 1) samples to test their catalytic performance, shown in Figure 7. No matter the value of x, CuxCo1–xMoO4/g-C3N4 samples were active in the hydrolysis of ammonia borane. CuMoO4/g-C3N4, without Co element, showed the lowest catalytic activity with a STY value of 11.7 m o l H 2 m o l c a t 1 m i n 1 . When Co element was present, the activities increased remarkably, and the best molar ratio of Cu to Co was 4:6 with a STY value of 75.7 m o l H 2 m o l c a t 1 m i n 1 . It is worth noting that the STY of CoMoO4/g-C3N4 was 35.35 m o l H 2 m o l c a t 1 m i n 1 , which is lower than those of Cu0.2Co0.8MoO4/g-C3N4, Cu0.4Co0.6MoO4/g-C3N4, and Cu0.6Co0.4MoO4/g-C3N4, indicating that the coexistence of cobalt and copper in the sample is helpful to improving the catalytic activity. Notably, when CoMoO4 served as a catalyst, there was no hydrogen released from the AB solution within the first minute. After then, hydrogen was constantly produced. This observation indicates that there is an induction period for the CoMoO4 catalyst. In contrast, all the Cu-containing catalysts in this study had no induction period in AB hydrolysis, illustrating that copper element can shorten the induction period of catalysis. These conclusions are in a good agreement with those in previous reports [22,32]. We also synthesized CoMoO4 and CuMoO4, which had STY values of 25.5 and 3.9 m o l H 2 m o l c a t 1 m i n 1 , respectively; these STY values are lower than those of CoMoO4/g-C3N4 and CuMoO4/g-C3N4, further confirming that the introduction of g-C3N4 enhanced the catalytic activity.
To understand the relationship between catalyst dosage and hydrogen production rate, the hydrogen evolution curves for different amounts of Cu0.4Co0.6MoO4/g-C3N4 are shown in Figure 8. As the amount of Cu0.4Co0.6MoO4/g-C3N4 increased, the rate of hydrogen production also increased. By fitting the curve of ln(catalyst) vs. ln(rate), the relationship between the catalyst dosage and the hydrogen production rate was clarified. As shown in Figure 8b, the slope of the fitted curve was 0.901, indicating that the hydrolytic process is a pseudo-first-order reaction in the initial stage. The hydrogen production rates of the catalysts decreased gradually due to the diffusion rate limitation of AB and catalyst poisoning. This result is consistent with those in previous literature [5]. Therefore, hydrogen production can be controlled easily by adjusting the catalyst dosage in practice.
Figure 9a shows the relationship between the amount of ammonia borane and the volume of hydrogen production. It is interesting to note that the hydrogen release rates were almost the same at the initial stage of the hydrolytic reaction, demonstrating that the AB concentration has no pronounced effect on the initial reaction rate of hydrogen production. As shown in Figure 9b, the slope of ln(rate) vs. ln(NH3BH3) was 0.0222, which is close to zero, indicating that the catalytic reaction rate is independent of the AB concentration due to a zero-order reaction.
As we know, the temperature exerts a significant influence on catalytic activity of a catalyst. We also explored the relationship between the rate of hydrogen generation and catalytic temperature in the range 298 to 313 K (Figure 10). With increasing temperature, the slope of the hydrogen evolution curve became larger, and the time taken for complete hydrogen release became shorter. According to the Arrhenius formula, the apparent activation energy (Ea) can be calculated after fitting the curve of ln k vs. 1/T, where the k value is the slope of hydrogen generation vs. time. The Ea value of Cu0.4Co0.6MoO4/g-C3N4 was as low as 14.46 kJ mol−1. In general, Ea can be used to roughly assess the catalytic activity of a catalyst. The low activation energy indicates that its energy barrier is low in the catalytic process, hinting that the catalysts possess high catalytic activity. Table 1 lists the STY values and activation energy data of some representative noble-metal-free catalysts in the literature and those of our Cu0.4Co0.6MoO4/g-C3N4. As can be seen, the STY value of Cu0.4Co0.6MoO4/g-C3N4 is higher than those of most noble-metal-free catalysts in the literature. Therefore, Cu0.4Co0.6MoO4/g-C3N4 is one of the best potential catalysts for AB hydrolysis.

3. Materials and Methods

3.1. Synthesis of Catalysts

All chemical reagents were analytically pure and had not been further purified. Deionized water was used during the experiment. To synthesize g-C3N4, 25 g urea was calcined at 550 °C for 6 hours in air with a heating rate of 2.5 °C min−1. A resultant yellow soft and porous powder was obtained. For the purpose of synthesizing 10 wt % CuxCo1–xMoO4/g-C3N4, 0.4 g g-C3N4 was dispersed in 40 mL ethanol and 20 mL deionized water. Then, 0.18 mmol copper chloride and cobalt chloride mixture solution was dropped into the g-C3N4 suspension under magnetic stirring for 2 hours. After that, 0.18 mmol sodium molybdate was added into the mixture solution. The molar ratios of Cu to Co to Mo were set at x:(1–x):1 (x = 0, 0.2, 0.4, 0.6, 0.8, 1). Then, 0.1 M ammonium hydroxide was used to adjust the pH of the suspension to 7. Finally, the suspension was transferred into a Teflon-lined stainless reaction vessel, which was placed in an oven and heated at 160 °C for 8 hours. The obtained samples were centrifugally separated, washed, dried, and finally calcined at 350 °C for 2 hours. For comparison, unsupported Cu0.4Co0.6MoO4 was also prepared in a similar process to that described above, except that no g-C3N4 was used.

3.2. Characterizations

Powder X-ray diffraction (XRD) was measured to obtain crystallographic information using a Bruker D8 Discover X-ray diffractometer with Cu Kα radiation (λ = 1.5406Å) in the range of 2θ = 10°–80° (Bruker, Billerica, MA, USA). The Fourier transform infrared spectrum (FTIR) was measured to acquire functional group information using a Nicolet 6700 FTIR spectrometer in the range of 450–3000 cm−1 (ThermoFisher Scientific, Waltham, MA, USA). Microstructures and the EDS spectrum were analyzed using a JEOL-7100F field-emission scanning electron microscope (FESEM) (Jeol Ltd., Akishima, Japan). The elements and oxidation states were analyzed using a Kratos Axis Ultra DLD X-ray photoelectron spectrometer (XPS) (Kratos Analytical, New York, NY, USA). The molar ratios of Cu, Co, and Mo in different samples were determined using an Agilent 7800 inductively coupled plasma mass spectrometer (ICP-MASS) (Agilent Technologies Inc., Santa Clara, CA, USA).

3.3. Catalytic Experiments

Due to the significant influence of temperature on catalytic performance, the catalyst properties were tested at 298 K. Typically, 5.0 mg of active substance, was dispersed in 5 mL deionized water. Then, 15 mL of mixture solution of 3 mmol ammonia borane and 20 mmol NaOH was quickly added to the CuxCo1–xMoO4/g-C3N4 solution. The volume of the generated hydrogen was measured by a drainage method, and the hydrogen production was recorded at intervals.

4. Conclusions

In summary, we synthesized a series of CuxCo1–xMoO4/g-C3N4 (x = 0, 0.2, 0.4, 0.6, 0.8, 1) and unsupported Cu0.4Co0.6MoO4 for the hydrolysis of ammonia borane. As a representative catalyst, Cu0.4Co0.6MoO4/g-C3N4 was characterized by XRD, FTIR, SEM, and XPS. The catalytic activity of Cu0.4Co0.6MoO4/g-C3N4 was measured under different temperatures, different catalyst dosages, and different amounts of ammonia borane. It was found that Cu0.4Co0.6MoO4/g-C3N4 showed the best catalytic activity with a STY value of 75.7 m o l H 2 m o l c a t 1 m i n 1 . The well-controlled composition and the well-dispersed active substances contributed to the improvement of the catalytic performance.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/9/9/714/s1, Figure S1: XRD patterns of CuxCo1–xMoO4/g-C3N4 (x = 0, 0.2, 0.6, 0.8, 1) and Cu0.4Co0.6MoO4, Figure S2: FTIR pattern of CuxCo1–xMoO4/g-C3N4, Figure S3: EDS spectra of CuxCo1–xMoO4/g-C3N4 for (a) x = 0, (b) x = 0.2, (c) x = 0.6, (d) x = 0.8, and (e) x = 1 and (f) Cu0.4Co0.6MoO4, Table S1: Element ratios of Cu, Co, and Mo in CuxCo1–xMoO4/g-C3N4.

Author Contributions

Sample synthesis and manuscript writing, J.L.; investigation of the catalytic performance, F.L.; characterization and analysis of the sample, J.L.; funding acquisition and reviewing the manuscript, H.L. and Q.L.

Funding

This work was funded by the National Natural Science Foundation of China (No. 21606050, U1801257), the Natural Science Foundation of Guangdong Province (No. 2018A030313859), Pearl River Science and Technology New Star Project (No. 201806010039), the Major Project of Fundamental and Application Research of the Department of Education of Guangdong Province (No. 2017KZDXM079), the Science & Technology project of Huizhou City (No. 2017C0412028), and the Natural Science Foundation of Huizhou University (No. 20180927172750326).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sartbaeva, A.; Kuznetsov, V.L.; Wells, S.A.; Edwards, P.P. Hydrogen nexus in a sustainable energy future. Energy Environ. Sci. 2008, 1, 79. [Google Scholar] [CrossRef]
  2. Schlapbach, L.; Züttel, A. Hydrogen-storage materials for mobile applications. Nature 2001, 414, 353–358. [Google Scholar] [CrossRef] [PubMed]
  3. Jacobson, M.Z.; Colella, W.G.; Golden, D.M. Cleaning the air and improving health with hydrogen fuel-cell vehicles. Science 2005, 308, 1901–1905. [Google Scholar] [CrossRef] [PubMed]
  4. Liu, C.; Fan, Y.Y.; Liu, M.; Cong, H.T.; Cheng, H.M.; Dresselhaus, M.S. Hydrogen Storage in Single-Walled Carbon Nanotubes at Room Temperature. Science 1999, 286, 1127–1129. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Peng, C.Y.; Kang, L.; Cao, S.; Chen, Y.; Lin, Z.S.; Fu, W.F. Nanostructured Ni2P as a Robust Catalyst for the Hydrolytic Dehydrogenation of Ammonia-Borane. Angew. Chem. Int. Ed. Engl. 2015, 54, 15725–15729. [Google Scholar] [CrossRef] [PubMed]
  6. Gao, M.; Yu, Y.; Yang, W.; Li, J.; Xu, S.; Feng, M.; Li, H. Ni nanoparticles supported on graphitic carbon nitride as visible light catalysts for hydrolytic dehydrogenation of ammonia borane. Nanoscale 2019, 11, 3506–3513. [Google Scholar] [CrossRef] [PubMed]
  7. Peng, B.; Chen, J. Ammonia borane as an efficient and lightweight hydrogen storage medium. Energy Environ. Sci. 2008, 1, 479–483. [Google Scholar] [CrossRef]
  8. Wen, M.; Cui, Y.; Kuwahara, Y.; Mori, K.; Yamashita, H. Non-Noble-Metal Nanoparticle Supported on Metal-Organic Framework as an Efficient and Durable Catalyst for Promoting H2 Production from Ammonia Borane under Visible Light Irradiation. ACS Appl. Mater. Interfaces 2016, 8, 21278–21284. [Google Scholar] [CrossRef] [PubMed]
  9. Lu, D.; Liao, J.; Zhong, S.; Leng, Y.; Ji, S.; Wang, H.; Wang, R.; Li, H. Cu0.6Ni0.4Co2O4 nanowires, a novel noble-metal-free catalyst with ultrahigh catalytic activity towards the hydrolysis of ammonia borane for hydrogen production. Int. J. Hydrog. Energy 2018, 43, 5541–5550. [Google Scholar] [CrossRef]
  10. Navlani-García, M.; Mori, K.; Nozaki, A.; Kuwahara, Y.; Yamashita, H. Highly efficient Ru/carbon catalysts prepared by pyrolysis of supported Ru complex towards the hydrogen production from ammonia borane. Appl. Catal. A Gen. 2016, 527, 45–52. [Google Scholar] [CrossRef]
  11. Chandra, M.; Xu, Q. Room temperature hydrogen generation from aqueous ammonia-borane using noble metal nano-clusters as highly active catalysts. J. Power Sources 2007, 168, 135–142. [Google Scholar] [CrossRef]
  12. Jiang, H.L.; Umegaki, T.; Akita, T.; Zhang, X.B.; Haruta, M.; Xu, Q. Bimetallic Au–Ni nanoparticles embedded in SiO2 nanospheres: Synergetic catalysis in hydrolytic dehydrogenation of ammonia borane. Chem. Eur. J. 2010, 16, 3132–3137. [Google Scholar] [CrossRef] [PubMed]
  13. Chen, G.; Desinan, S.; Rosei, R.; Rosei, F.; Ma, D. Synthesis of Ni–Ru alloy nanoparticles and their high catalytic activity in dehydrogenation of ammonia borane. Chem. Eur. J. 2012, 18, 7925–7930. [Google Scholar] [CrossRef] [PubMed]
  14. Zheng, Y.; Jiao, Y.; Zhu, Y.; Cai, Q.; Vasileff, A.; Li, L.H.; Han, Y.; Chen, Y.; Qiao, S. Molecule-level g-C3N4 coordinated transition metals as a new class of electrocatalysts for oxygen electrode reactions. J. Am. Chem. Soc. 2017, 139, 3336–3339. [Google Scholar] [CrossRef] [PubMed]
  15. Li, X.; Wang, X.; Antonietti, M. Solvent-free and metal-free oxidation of toluene using O2 and g-C3N4 with nanopores: Nanostructure boosts the catalytic selectivity. ACS Catal. 2012, 10, 2082–2086. [Google Scholar] [CrossRef]
  16. Habibi-Yangjeh, A.; Mousavi, M.; Nakata, K. Boosting visible-light photocatalytic performance of g-C3N4/Fe3O4 anchored with CoMoO4 nanoparticles: Novel magnetically recoverable photocatalysts. J. Photochem. Photobiol. A Chem. 2019, 368, 120–136. [Google Scholar] [CrossRef]
  17. Guo, L.; Cai, Y.; Ge, J.; Zhang, Y.; Gong, L.; Li, X.; Wang, K.; Ren, Q.; Su, J.; Chen, J. Multifunctional Au-Co@CN Nanocatalyst for Highly Efficient Hydrolysis of Ammonia Borane. ACS Catal. 2015, 5, 388–392. [Google Scholar] [CrossRef]
  18. Lu, R.; Hu, M.; Xu, C.; Wang, Y.; Zhang, Y.; Xu, B.; Gao, D.; Bi, J.; Fan, G. Hydrogen evolution from hydrolysis of ammonia borane catalyzed by Rh/g-C3N4 under mild conditions. Int. J. Hydrog. Energy 2018, 43, 7038–7045. [Google Scholar] [CrossRef]
  19. Kahri, H.; Sevim, M.; Metin, Ö. Enhanced catalytic activity of monodispersed AgPd alloy nanoparticles assembled on mesoporous graphitic carbon nitride for the hydrolytic dehydrogenation of ammonia borane under sunlight. Nano Res. 2016, 10, 1627–1640. [Google Scholar] [CrossRef]
  20. Jia, H.; Chen, X.; Song, X.; Zheng, X.; Guan, X.; Liu, P. Graphitic carbon nitride-chitosan composites–anchored palladium nanoparticles as high-performance catalyst for ammonia borane hydrolysis. Int. J. Energy Res. 2018, 43, 535–543. [Google Scholar] [CrossRef]
  21. Wang, Q.; Xu, C.; Ming, M.; Yang, Y.; Xu, B.; Wang, Y.; Zhang, Y.; Wu, J.; Fan, G. In Situ Formation of AgCo Stabilized on Graphitic Carbon Nitride and Concomitant Hydrolysis of Ammonia Borane to Hydrogen. Nanomaterials 2018, 8, 280. [Google Scholar] [CrossRef]
  22. Lu, D.; Li, J.; Lin, C.; Liao, J.; Feng, Y.; Ding, Z.; Li, Z.; Liu, Q.; Li, H. A Simple and Scalable Route to Synthesize CoxCu1- xCo2O4@CoyCu1- yCo2O4 Yolk-Shell Microspheres, A High-Performance Catalyst to Hydrolyze Ammonia Borane for Hydrogen Production. Small 2019, 15, e1805460. [Google Scholar] [CrossRef] [PubMed]
  23. Wang, S.; Li, D.; Sun, C.; Yang, S.; Guan, Y.; He, H. Synthesis and characterization of g-C3N4/Ag3VO4 composites with significantly enhanced visible-light photocatalytic activity for triphenylmethane dye degradation. Appl. Catal. B Environ. 2014, 144, 885–892. [Google Scholar] [CrossRef]
  24. Cheng, J.; Huang, W. Effect of cobalt (nickel) content on the catalytic performance of molybdenum carbides in dry-methane reforming. Fuel Process. Technol. 2010, 91, 185–193. [Google Scholar] [CrossRef]
  25. Amanulla, A.M.; Shahina, S.K.J.; Sundaram, R.; Magdalane, C.M.; Kaviyarasu, K.; Letsholathebe, D.; Mohamed, S.B.; Kennedy, J.; Maaza, M. Antibacterial, magnetic, optical and humidity sensor studies of beta-CoMoO4-Co3O4 nanocomposites and its synthesis and characterization. J. Photochem. Photobiol. B 2018, 183, 233–241. [Google Scholar] [CrossRef] [PubMed]
  26. Tan, H.; Chen, W.; Liu, D.; Feng, X.; Li, Y.; Yan, A.; Wang, E. Two diphosphonate-functionalized asymmetric polyoxomolybdates with catalytic activity for oxidation of benzyl alcohol to benzaldehyde. Dalton Trans. 2011, 40, 8414–8418. [Google Scholar] [CrossRef] [PubMed]
  27. Li, C.; Du, Y.; Wang, D.; Yin, S.; Tu, W.; Chen, Z.; Kraft, M.; Chen, G.; Xu, R. Unique P–Co–N Surface Bonding States Constructed on g-C3N4 Nanosheets for Drastically Enhanced Photocatalytic Activity of H2 Evolution. Adv. Funct. Mater. 2017, 27, 1604328. [Google Scholar] [CrossRef]
  28. Li, Q.; Li, Y.; Zhao, J.; Zhao, S.; Zhou, J.; Chen, C.; Tao, K.; Liu, R.; Han, L. Ultrathin nanosheet-assembled hollow microplate CoMoO4 array derived from metal-organic framework for supercapacitor with ultrahigh areal capacitance. J. Power Sources 2019, 430, 51–59. [Google Scholar] [CrossRef]
  29. Yu, M.Q.; Jiang, L.X.; Yang, H.G. Ultrathin nanosheets constructed CoMoO4 porous flowers with high activity for electrocatalytic oxygen evolution. Chem. Commun. 2015, 51, 14361–14364. [Google Scholar] [CrossRef] [PubMed]
  30. Yu, B.; Jiang, G.; Xu, W.; Cao, C.; Liu, Y.; Lei, N.; Evariste, U.; Ma, P. Construction of NiMoO4/CoMoO4 nanorod arrays wrapped by Ni-Co-S nanosheets on carbon cloth as high performance electrode for supercapacitor. J. Alloys Compd. 2019, 799, 415–424. [Google Scholar] [CrossRef]
  31. Fernandes, R.; Patel, N.; Miotello, A.; Calliar, L. Co–Mo–B–P Alloy with Enhanced Catalytic Properties for H2 Production by Hydrolysis of Ammonia Borane. Top. Catal. 2012, 55, 1032–1039. [Google Scholar] [CrossRef]
  32. Liao, J.; Lu, D.; Diao, G.; Zhang, X.; Zhao, M.; Li, H. Co0.8Cu0.2MoO4 Microspheres Composed of Nanoplatelets as a Robust Catalyst for the Hydrolysis of Ammonia Borane. ACS Sustain. Chem. Eng. 2018, 6, 5843–5851. [Google Scholar] [CrossRef]
  33. Wang, C.; Tuninetti, J.; Wang, Z.; Zhang, C.; Ciganda, R.; Salmon, L.; Moya, S.; Ruiz, J.; Astruc, D. Hydrolysis of ammonia-borane over Ni/ZIF-8 nanocatalyst: High efficiency, mechanism and controlled hydrogen release. J. Am. Chem. Soc. 2017, 139, 11610. [Google Scholar] [CrossRef] [PubMed]
  34. Zheng, H.; Feng, K.; Shang, Y.; Kang, Z.; Sun, X.; Zhong, J. Cube-Like CuCoO Nanostructures on Reduced Graphene Oxide for H2 Generation from Ammonia Borane. Inorg. Chem. Front. 2018, 5, 1180. [Google Scholar] [CrossRef]
  35. Fu, Z.C.; Xu, Y.; Chan, S.L.; Wang, W.W.; Li, F.; Liang, F.; Chen, Y.; Lin, Z.S.; Fu, W.F.; Che, C.M. Highly efficient hydrolysis of ammonia borane by anion (OH, F, Cl)-tuned interactions between reactant molecules and CoP nanoparticles. Chem. Commun. 2017, 53, 705–708. [Google Scholar] [CrossRef] [PubMed]
  36. Feng, K.; Zhong, J.; Zhao, B.; Zhang, H.; Xu, L.; Sun, X.; Lee, S.T. CuxCo1–xO Nanoparticles on Graphene Oxide as A Synergistic Catalyst for High-Efficiency Hydrolysis of Ammonia–Borane. Angew. Chem. Int. Ed. Engl. 2016, 55, 11950–11954. [Google Scholar] [CrossRef] [PubMed]
  37. Yao, Q.; Lu, Z.-H.; Huang, W.; Chen, X.; Zhu, J. High Pt-like activity of the Ni–Mo/graphene catalyst for hydrogen evolution from hydrolysis of ammonia borane. J. Mater. Chem. A 2016, 4, 8579–8583. [Google Scholar] [CrossRef]
  38. Bulut, A.; Yurderi, M.; Ertas, İ.E.; Celebi, M.; Kaya, M.; Zahmakiran, M. Carbon dispersed copper-cobalt alloy nanoparticles: A cost-effective heterogeneous catalyst with exceptional performance in the hydrolytic dehydrogenation of ammonia-borane. Appl. Catal. B Environ. 2016, 180, 121–129. [Google Scholar] [CrossRef]
  39. Zhang, J.; Li, H.; Zhang, H.; Zhu, Y.; Mi, G. Porously hierarchical Cu@Ni cubic-cage microstructure: Very active and durable catalyst for hydrolytically liberating H2 gas from ammonia borane. Renew. Energy 2016, 99, 1038–1045. [Google Scholar] [CrossRef]
  40. Song, F.Z.; Zhu, Q.L.; Yang, X.C.; Xu, Q. Monodispersed CuCo Nanoparticles Supported on DiamineFunctionalized Graphene as a Non-noble Metal Catalyst for Hydrolytic Dehydrogenation of Ammonia Borane. ChemNanoMat 2016, 2, 942–945. [Google Scholar] [CrossRef]
  41. Du, X.; Yang, C.; Zeng, X.; Wu, T.; Zhou, Y.; Cai, P.; Cheng, G.; Luo, W. Amorphous NiP supported on rGO for superior hydrogen generation from hydrolysis of ammonia borane. Int. J. Hydrog. Energy 2017, 42, 14181–14187. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of graphitic carbon nitride (g-C3N4) and Cu0.4Co0.6MoO4/g-C3N4 (a) and enlarged XRD pattern of Cu0.4Co0.6MoO4/g-C3N4 (b).
Figure 1. XRD patterns of graphitic carbon nitride (g-C3N4) and Cu0.4Co0.6MoO4/g-C3N4 (a) and enlarged XRD pattern of Cu0.4Co0.6MoO4/g-C3N4 (b).
Catalysts 09 00714 g001
Figure 2. FTIR spectra of Cu0.4Co0.6MoO4, g-C3N4, and Cu0.4Co0.6MoO4/g-C3N4.
Figure 2. FTIR spectra of Cu0.4Co0.6MoO4, g-C3N4, and Cu0.4Co0.6MoO4/g-C3N4.
Catalysts 09 00714 g002
Figure 3. FESEM images of (a) g-C3N4, (b,c) Cu0.4Co0.6MoO4, and (d,e) Cu0.4Co0.6MoO4/g-C3N4; (f) EDS spectrum of point analysis.
Figure 3. FESEM images of (a) g-C3N4, (b,c) Cu0.4Co0.6MoO4, and (d,e) Cu0.4Co0.6MoO4/g-C3N4; (f) EDS spectrum of point analysis.
Catalysts 09 00714 g003
Figure 4. XPS pattern of Cu0.4Co0.6MoO4/g-C3N4: (a) C 1s; (b) N 1s; (c) Co 2p; (d) Cu 2p; (e) Mo 3d; (f) O 1s.
Figure 4. XPS pattern of Cu0.4Co0.6MoO4/g-C3N4: (a) C 1s; (b) N 1s; (c) Co 2p; (d) Cu 2p; (e) Mo 3d; (f) O 1s.
Catalysts 09 00714 g004
Figure 5. Time vs. n(H2)/n(AB) during the catalytic hydrolysis of ammonia borane (AB) with catalysts Cu0.4Co0.6MoO4, g-C3N4 and Cu0.4Co0.6MoO4/g-C3N4.
Figure 5. Time vs. n(H2)/n(AB) during the catalytic hydrolysis of ammonia borane (AB) with catalysts Cu0.4Co0.6MoO4, g-C3N4 and Cu0.4Co0.6MoO4/g-C3N4.
Catalysts 09 00714 g005
Figure 6. Scheme of the catalytic properties of bulk Cu0.4Co0.6MoO4 and Cu0.4Co0.6MoO4/g-C3N4.
Figure 6. Scheme of the catalytic properties of bulk Cu0.4Co0.6MoO4 and Cu0.4Co0.6MoO4/g-C3N4.
Catalysts 09 00714 g006
Figure 7. (a) Time vs. n(H2)/n(AB) during the catalytic hydrolysis of AB with catalyst CuxCo1–xMoO4/g-C3N4; (b) STY values with x for CuxCo1–xMoO4/g-C3N4.
Figure 7. (a) Time vs. n(H2)/n(AB) during the catalytic hydrolysis of AB with catalyst CuxCo1–xMoO4/g-C3N4; (b) STY values with x for CuxCo1–xMoO4/g-C3N4.
Catalysts 09 00714 g007
Figure 8. (a) n(H2)/n(AB) vs. reaction time at different Cu0.4Co0.6MoO4/g-C3N4 dosages; (b) the relationship between the logarithms of the catalyst mass and the catalytic rate.
Figure 8. (a) n(H2)/n(AB) vs. reaction time at different Cu0.4Co0.6MoO4/g-C3N4 dosages; (b) the relationship between the logarithms of the catalyst mass and the catalytic rate.
Catalysts 09 00714 g008
Figure 9. (a) Volume of hydrogen vs. reaction time at different Cu0.4Co0.6MoO4/g-C3N4 dosages; (b) the relationship between the logarithms of the catalyst mass and the catalytic rate.
Figure 9. (a) Volume of hydrogen vs. reaction time at different Cu0.4Co0.6MoO4/g-C3N4 dosages; (b) the relationship between the logarithms of the catalyst mass and the catalytic rate.
Catalysts 09 00714 g009
Figure 10. (a) Hydrogen evolution at different reaction temperatures in the range of 298–313 K; (b) the relationship between logarithm of k and 1/T.
Figure 10. (a) Hydrogen evolution at different reaction temperatures in the range of 298–313 K; (b) the relationship between logarithm of k and 1/T.
Catalysts 09 00714 g010
Table 1. Comparison of the STY value and apparent activation energy of our Cu0.4Co0.6MoO4/g-C3N4 catalyst with those of some other representative noble-metal-free catalysts in the literature.
Table 1. Comparison of the STY value and apparent activation energy of our Cu0.4Co0.6MoO4/g-C3N4 catalyst with those of some other representative noble-metal-free catalysts in the literature.
CatalystsSTY
(molH2 molcat−1 min−1)
STY without support
(molH2 molcat−1 min−1)
Ea
(kJ mol−1)
Reference
Ni/ZIF-885.7/28[33]
Cu0.5Co0.5O-rGO81.7/−45.26[34]
Cu0.4Co0.6MoO4/g-C3N475.717.614.46This work
CoP72.2//[35]
Cu0.8Co0.2O-GO70.0/45.53[36]
Ni0.9Mo0.1/graphene66.72.3/[37]
Co0.8Cu0.2MoO455//[32]
Cu0.49Co0.51/C45/51.9[38]
Cu0.8Ni0.241.9/40.53[39]
CuCo/graphene411254.89[40]
NiP40.4/44.6[5]
Ni91P9/rGO13.3/34.7[41]
Ni/g-C3N418.7/36[6]

Share and Cite

MDPI and ACS Style

Li, J.; Li, F.; Liao, J.; Liu, Q.; Li, H. Cu0.4Co0.6MoO4 Nanorods Supported on Graphitic Carbon Nitride as a Highly Active Catalyst for the Hydrolytic Dehydrogenation of Ammonia Borane. Catalysts 2019, 9, 714. https://doi.org/10.3390/catal9090714

AMA Style

Li J, Li F, Liao J, Liu Q, Li H. Cu0.4Co0.6MoO4 Nanorods Supported on Graphitic Carbon Nitride as a Highly Active Catalyst for the Hydrolytic Dehydrogenation of Ammonia Borane. Catalysts. 2019; 9(9):714. https://doi.org/10.3390/catal9090714

Chicago/Turabian Style

Li, Junhao, Fangyuan Li, Jinyun Liao, Quanbing Liu, and Hao Li. 2019. "Cu0.4Co0.6MoO4 Nanorods Supported on Graphitic Carbon Nitride as a Highly Active Catalyst for the Hydrolytic Dehydrogenation of Ammonia Borane" Catalysts 9, no. 9: 714. https://doi.org/10.3390/catal9090714

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