Samikannu
Prabu
and
Kung-Yuh
Chiang
*
Graduate Institute of Environmental Engineering, National Central University, Tao-Yuan city, 32001, Taiwan. E-mail: kychiang@ncu.edu.tw
First published on 28th July 2020
The development of high catalytic effective catalysts for hydrogen generation through dehydrogenation (hydrolysis) of aqueous NH3BH3 (AB) solution is discussed in this work. Bimetallic NiPt, CoPt and monometallic Ni, Co, Pt nanoparticles (NPs) supported on mixed graphene oxide (GO) and reduced graphene oxide rGO (carbon materials) were produced and studied for hydrogen generation from AB hydrolytic dehydrogenation. Herein, we have developed fine, spherically-shaped bimetallic and monometallic Ni, Co, Pt NPs on mixed GO and rGO for extremely high reactant productivity in aqueous AB hydrolysis. The Co0.8Pt0.2/GO and rGO, Ni0.8Pt0.2/GO and rGO catalysts show high catalytic performance and high turnover frequency (TOF) of 230.76 and 214.28 (H2) mol. (cat. metal) mol−1 min−1 at 25 °C. This is the greatest efficiency ever shown for transition metal-doped GO and rGO catalysts. The catalysts additionally show superior catalytic stability by maintaining up to 98% activity after 7 runs at 25 °C. The development of well-organized and inexpensive Co0.8Pt0.2/GO and rGO, Ni0.8Pt0.2/GO and rGO catalysts improve the possibility of using aqueous AB as chemical hydrogen storage. This permits the discovery of additional hydrogen fuel-cell applications. The simple and facile production of other GO and rGOs can assist in transition metal NPs.
Some transition metal NPs have shown amazing catalytic efficiencies.5,6 Therefore, studying the reasonable relationship between the catalytic performance and the metal support is essential. The quantitative recognition of the nature of metal–support interfaces in the absence of detailed information on hydrolysis reaction and catalytic active sites remains a field of interest. The established hydrolysis reaction methods under homogeneous catalysis cannot be straightforwardly connected to heterogeneous catalysis in the presence of the phase interface and complex active sites in heterogeneous catalysis.7
Hydrogen generation by means of thermal decomposition, for example, aqueous AB solution hydrolysis in the presence of an acceptable catalyst produces three moles of hydrogen gas per mole of AB solution at 25 °C (eqn (1)). This makes it a compelling catalyst for hydrogen generation from AB.
NH3BH3(aq) + 2H2O (l) → NH4 + (aq) + BO2 − (aq) + 3H2 (g) | (1) |
For the most part, Pt,8–12 Ru,13–15 and Pd16 based catalysts outshine non-noble metals, such as Ni, Cr,17 Co,18–20 RuPd@GO,21 Ag/Pd,22,23 Fe,24,25 Ag–Ni based nanoparticles,26 Au–Pd,27 mesoporous carbon nitride supported Pd and Pd–Ni NPs,28 PdNi–CeO2,29 NiPt NPs with supported CeO230 (efficient hydrogen generation from an alkaline solution of hydrazine), N-doped graphene supported Co–CeOx,31 and transition metal nanoparticles with GO32 for hydrogen generation under the above conditions. Though there has been development in the catalytic efficiencies of bimetallic catalysts, for example, Co–Pd33,34 and core–shell species such as Au@Co,35 and monometallic catalysts, the general utilization of these noble metals is restricted by their high cost and the extraordinary amount required. To solve this issue, the search for abundant metal catalysts with durable catalytic activities that can be utilized at 25 °C is very urgent and essential.
One useful surface approach is to modulate the contribution of active metal NPs. A single-atom two-dimensional material exhibits interesting properties, such as a high specific surface area, high catalytic efficiency, fine and spherically-shaped particles, and conserved charge transfer, and is, therefore, an ideal substrate for the advancement and affixing of metal NPs such as, graphene, GO, and carbon nanotubes.36,37 In spite of the staggering expense of graphene, graphene-supported metal NPs have become noteworthy supports because of their potential applications in a few specialized areas such as catalysis, devices, and energy conversion.38 Of late, metal oxides such as TiO2, SnO2, Fe3O4, SiO2, and CeO2 have been commonly utilized as catalyst promoters to build reactant stability and metal nanocatalyst activity.39 Among the examined metal oxides, the rare-earth metal oxide CeO2 is of considerable importance because of its rich oxygen defects, high oxygen storage capacity, and cost-effectiveness.40–45 In this situation, the presence of an M/CeO2/graphene triple convergence with the CeO2 mixture, a metal, and graphene may potentially produce a catalyst with remarkably improved catalytic activity for the dehydrogenation of aqueous AB as well as dispensability in the catalytic process.
Herein, we report a green and facile preparation for the synthesis of well-dispersed Co0.8Pt0.2/GO and rGO, Ni0.8Pt0.2/GO rGO catalysts that were utilized for catalytic hydrogen generation from AB hydrolytic dehydrogenation at 25 °C. Incredibly, the prepared noble-metal Co0.8Pt0.2/GO and rGO, Ni0.8Pt0.2/GO and rGO catalysts demonstrate higher catalytic performance, 100% hydrogen selectivity, and strong durability for hydrogen generation from AB, with 3 equivalents per mmole of AB generated within 0.6 and 0.5 min, respectively. The CoPt/GO and rGO, NiPt/GO and rGO catalysts demonstrate that they have the most important role in improving the metal NP catalytic efficiency. Because of the synergistic effect among CoPt and NiPt NPs supported on GO and rGO and the strong metal–support interface between the metals and the supporter, the prepared Co0.8Pt0.2/GO and rGO, Ni0.8Pt0.2/GO and rGO catalysts showed greater catalytic efficiency for hydrogen generation from aqueous AB solution, with TOF of 230.76 and 214.28 min−1 at 25 °C, respectively.
The above method was also utilized for characterizing the catalytic efficiency of other catalysts for H2 generation from aqueous AB solution hydrolysis. The nM/AB molar ratio for each one of the reactions was maintained at 0.04 M. The improvement in the catalytic performance over the Ni0.8Pt0.2/GO and rGO (optimum catalyst), Co0.8Pt0.2/GO and rGO NPs was observed at various temperatures (278 K, 283 K, 288 K, 293 K, 298 K, and 303 K) to determine the actuation energy (Ea) for the aqueous AB hydrolysis reaction. An equivalent procedure was additionally helpful to for determining the catalytic performance of the catalytic reaction with 100 mg of nanoparticles as the catalyst in the presence of different added substances (NaOH, KOH, NaH2PO4, NH4OH, and NaCl). However, the above-mentioned substances were added prior to adding the aqueous AB solutions into the reaction flask.
TOF = nH2/(nM × t) | (S1) |
The Ni0.8Pt0.2/GO and rGO, Co0.8Pt0.2/GO and rGO NPs morphologies were characterized using scanning electron microscopy (FESEM) and transition electron microscopy (TEM) images, as shown in Fig. 1. The process can be seen in Fig. 1b; NiPt NPs with fine and small sizes of about 1.5 nm (Fig. 1c) are well-spread onto the GO and rGO sheets, as shown by the FESEM result (Fig. 1a). In any case, the synthesized CoPt/GO and rGO NPs have a small, spherically-shaped particle size of about 1.9 nm (Fig. 1b′ and c′). By observing the TEM images of Ni0.8Pt0.2/GO and rGO NPs, Co0.8Pt0.2/GO and rGO NPs, it can be clearly seen that CoO doping into the NiO NPs can cause a decrease in the metal NP sizes. In order to further investigate, the metal-doped GO NPs obtained with a different (Co + Ni) molar ratio are named as CoNi/GO. The typical FESEM and TEM images of the obtained CoNi/GO catalysts are demonstrated in Fig. S3 (ESI†). The nanosheet shape of GO is wide-ranging and uniform after the CoNi NPs are loaded via the reduction reaction. It can be seen from the FESEM (Fig. S3a, ESI†) and bright-field TEM (Fig. S3b, ESI†) images that the CoNi NPs acquired through the rapid reduction process (Co + Ni) are highly distributed into the Co0.6Ni0.4/GO and Co0.4Ni0.6/GO NPs. The mean particle sizes are about 3.2 nm and 4.2 nm (Fig. S3c and f, ESI†), respectively, with further evidence of the existence of Co, Ni, and Pt in the doped samples. EDX was used to observe the elemental composition of the doped samples. After the EDX spectroscopy investigation, the results obtained reveal that the samples are composed of C, O, Co, Ni, and Pt elements that are found in all the catalysts and the total compositional atomic percentages of C, O, Co, Ni, and Pt were estimated. Table summarizes (inside the EDX spectrum) the elements in the nanostructures (Fig. S4, ESI†) and the occurrence of Ni, Co, C, and O was recognized by elemental mapping (Fig. S5 and S6, ESI†). It can be obviously seen that the distribution of Ni, Co, C, and O elements is very consistent with that shown in the FESEM images.
For investigating the feasibility of the NPs, we observed and associated the monometallic Pt/GO and Co/GO catalyst microstructure in the as-prepared catalysts. The TEM images of monometallic Pt/GO and Co/GO show that separate Pt/GO and Co/GO NPs are homogeneously dispersed on GO with an average particle size of about 4.3 nm and 30.2 nm (Fig. S7, ESI†), respectively. The particle sizes of the Pt and Co NPs in this study are smaller. Histograms of the NPs were achieved by including at least 50 particles. The Pt and Co NPs size of the Pt/GO catalyst is about 4.3 nm, whereas the Co size of the Co/GO catalyst is distributed at about 30.2 nm. The mean particle size of Pt/GO is smaller than that of the Co/GO NPs.
Fig. 2 demonstrates the characteristic powder-XRD patterns of the as-obtained GO and rGO and the metal mixtures. All the peaks could be correlated with the face-centered cubic (fcc) structures. The first graphene oxide (GO) sample demonstrated a strong diffraction peak centered at 2θ = 13.01, corresponding to an interlayer spacing of about 0.76 nm, compared to the (001) reflection of graphene oxide, which is broader than that of pristine graphite. After surface doping, the metal-doped (Ni, Co, and Pt) GO revealed negligible C (001) peaks compared to GO. For the metal-doped materials after minor heat treatment, the GO peak widened and the rGO (2θ = 26.39) (022) peak plane with an interlayer spacing of about 0.34 nm was obtained, representing the reduction of graphene oxide to rGO. The GO peak broadens on exposure to high temperature. Furthermore, as observed from the XRD pattern in Fig. 2, the diffraction peaks at 2θ = 39.64°, 46.36°, and 67.55° could be indexed to the characteristic (111), (200), and (220) planes of the cubic crystalline structured Pt, respectively. The NiO and CoO (PDF#96-900-8619) peaks are very strong for Co support to that NiO, CoO is improved detached on the GO, and rGO support with smaller and fine particles demonstrating their high crystallinity. These observed results clearly confirm the formation of the Ni, Co, and Pt/GO NPs. Moreover, the wide peaks at 31.16°, 36.64°, and 59.14° were likewise observed for the three metal oxides, representing the formation of rGO and GO doped with metal NPs. Compared with the monometallic catalysts, the bimetallic catalysts were determined to undergo considerably more structural alteration and therefore, a greater number of active sites were created om them for the catalytic reactions.
Fig. 2 The typical XRD patterns of the as-prepared GO nanosheet (a), CoNi/GO and rGO, (b), CoOPt/GO and rGO (c), and NiPt/GO and rGO NPs. |
Raman spectroscopy was used to further distinguish the reduction of GO, with two prominent peaks at 1338.26 and 1594 cm−1 conforming to the D and G bands of the C atomic crystals, respectively. The peaks for the NiPt/GO and rGO, CoPt/GO and rGO catalysts were red-shifted to 1339.06, 1346.48 and 1598.74, 1596.86 cm−1 for the D and G bands, respectively, after 1 h reduction at 500 °C, as shown in Fig. 3. The D-peak represents the defects in the lattice of the C atom, while the G-peak in-plane stretching vibration represents the sp2 hybridization of the C atom. Furthermore, the D/G intensity ratio was significantly broadened after reduction, from 0.83 for GO to 0.84 and 0.84 for NiPt/GO and rGO, CoPt/GO and rGO after 1 h reduction, representing that large amounts of graphitic carbon were present in the samples, respectively. The red-shift of the G band and an increase in the D/G intensity ratio both indicate the increase in the sp2 carbon regions.
Fig. 4 Hydrogen productivity vs. reaction time for hydrogen release from an aqueous AB solution (100 mM, 10 mL) catalyzed by M/GO and rGO (M–Co, Ni, and Pt) at 298 K (nNi/nAB = 0.04). |
Fig. 5 The corresponding TOF values for the dehydrogenation of ammonia borane solution (100 mM, 10 mL) catalyzed by M/GO and rGO (M–Co, Ni, and Pt) at 298 K (nNi/nAB = 0.04). |
Various supports changed the structures, bringing about different catalytic performances. At that point, the absorbent inorganic catalysts, Al2O3 and SiO2, were utilized as supports for the synthesis of Ni and Co NP catalysts using a similar method to that for NiPt/GO and rGO. The TEM images show that the Ni and Co NPs in NiPt/Al2O3, CoNi/Al2O3, and NiPt/SiO2 likewise have a small size (Fig. S8, ESI†). However, they were greater than those of NiPt/GO and rGO. The synergistic catalysts in NiPt/Al2O3, CoNi/Al2O3, and NiPt/SiO2 showed lower reactant dehydrogenation than NiPt/GO and rGO (Fig. 6) but their catalytic performance was still higher than that of the other reported noble-metal free catalysts (Table S3, ESI†). These distinctive catalyst efficiencies may be due to various materials, among which the high specific surface area of the metal NPs and the solid adsorption of NH3BH3 into their pores led to the quick reactant dehydrogenation of AB. In addition, these consequences, including various types of supports, can be additionally attributed to the assumption of improved synergistic effect over amorphous metal NPs.
In addition, the detailed examination of the kinetic reaction for the aqueous AB hydrolysis reaction was undertaken by changing the number of catalysts with the similar amount of AB at room temperature (25 °C). Fig. 7A shows the influence of the amount of catalyst on the catalytic hydrogen generation reaction, where the amount of AB was kept at 2 mmol at 25 °C and the catalytic amounts were set at 20, 40, 60, 80, and 100 mg. We can see that the catalysts can effectively start the AB complex hydrolysis. The hydrogen generation efficiency was improved with increasing catalyst amount within the measured range. Fig. 7B demonstrates the plot of moles of hydrogen generated versus the catalyst amount on a logarithmic scale. The slope of 1.99 shows that the catalytic hydrolysis reaction of aqueous AB is first-order with respect to the catalyst amount and the observation is in acceptable understanding.
(2) |
The activation energy is accordingly calculated from the Arrhenius equation (eqn (3)) between lnk and 1/T as follows
(3) |
To demonstrate the temperature effect on aqueous AB hydrolytic dehydrogenation, the kinetics of Ni0.8Pt0.2/GO and rGO, Co0.8Pt0.2/GO and rGO catalyzed hydrogen generation were studied at different temperatures (25–60 °C). Fig. 8a and c shows the hydrolytic dehydrogenation amount of AB catalyzed by Ni0.8Pt0.2/GO and rGO, Co0.8Pt0.2/GO and rGO NPs. The reaction was carried out at increased temperature in the range of 25–60 °C. The rate of the reaction is significantly improved on increasing the temperature. As indicated by the Arrhenius plot in Fig. 8b and d, the achieved activation energy (Ea) of AB hydrolysis catalyzed by Ni0.8Pt0.2/GO and rGO, Co0.8Pt0.2/GO and rGO NPs is 23.94 kJ mol−1 and 23.60 kJ mol−1, respectively, which is lower than those of the recently reported noble-metal catalysts.31 As shown in Fig. S9 (ESI†), the hydrogen generation rate increases with increasing temperature. The catalytic reactions for hydrogen generation from aqueous AB solution were completed in 0.6, 0.5, 0.4, 0.3, 0.2 min (Ni0.8Pt0.2/GO and rGO), 0.62, 0.55, 0.42, 0.33, 0.21 min (Co0.8Pt0.2/GO and rGO), and 25, 30, 40, 50, 60 °C, whose corresponding TOF values are 214.0, 250.0, 333.3, 500.0 and 750.0 min−1 (Ni0.8Pt0.2/GO and rGO), and 176.4, 230.7, 300.0, 428.5, 681.8 min−1 (Co0.8Pt0.2/GO and rGO) (Fig. S9, ESI†).
Seeing the significance of the catalytic stability in practical application, the recyclability and stability of the Ni0.8Pt0.2/GO and rGO, Co0.8Pt0.2/GO and rGO NPs were studied at 25 °C by adding the same amount of the AB solution. When the first cycle was completed, as shown in Fig. 9, the hydrogen generation volume was unaltered after 10 cycles. The initial activity was maintained, showing its great recyclability. However, the hydrogen generation amount displayed a slight decrease. After the stability test, the Ni0.8Pt0.2/GO and rGO, Co0.8Pt0.2/GO and rGO NPs were then characterized by TEM. The TEM image of the used Ni0.8Pt0.2/GO and rGO, Co0.8Pt0.2/GO and rGO NP catalysts clearly showed that no great change could be recognized for the morphology of the reused catalysts. The CoPt and NiPt NPs in the GO support were still well-scattered without the presence of clusters. The effective catalytic performance and stability can be similarly obtained with smaller-sized CoPt and NiPt NPs. In addition, the pyrolytically-obtained GO layer could likewise help to assist the CoPt and NiPt NPs.
Fig. 9 Reusability of the Ni0.8Pt0.2/GO and rGO, Co0.8Pt0.2/GO and rGO nanocatalysts for AB hydrolysis for 10 successive cycles at 25 °C. |
In further analysis, as shown in Fig. S10 (ESI†), the morphology of Ni0.8Pt0.2/GO and rGO, Co0.8Pt0.2/GO and rGO was well preserved. Although some mixing was detected, which might be the reason for the slight decrease in the activity during the durability test, the TEM images of the recovered catalyst show no significant variation in the particles (Fig. S10c, d, ESI†). Moreover, there is a small change in the particle size and agglomeration is observed, signifying that the structure of the Ni0.8Pt0.2/GO and rGO, Co0.8Pt0.2/GO and rGO NPs remains unchanged after usage. The Ni0.8Pt0.2/GO and rGO, Co0.8Pt0.2/GO and rGO catalysts demonstrate very high catalytic stability and durability towards AB hydrolytic dehydrogenation. In addition, Fig. S11 (ESI†) shows the X-ray diffraction (XRD) pattern of the reused catalyst for Ni0.8Pt0.2/GO and rGO NPs. Obviously, in addition to the GO (2θ = 13.26°) diffraction peak, only one broad diffraction peak at about 2θ = 43.0°, representing the disappearance of rGO, was observed. This indicates that the Ni0.8Pt0.2/GO and rGO NPs have a low crystalline structure, as shown in Fig. S11 (ESI†).
The aqueous AB solution catalytic efficiency was confirmed and studied at various temperatures in the range of 5–15 °C starting with Ni0.8Pt0.2/GO and rGO and 100 mM AB in 10 mL of H2O. As shown in Fig. 10A, the hydrogen generation rate steadily increased with increasing temperature and the TOF value reached the maximum at 15 °C. Thus, the Ni0.8Pt0.2/GO and rGO catalyst was chosen in all the further experiments.
Based on the above presumptions and the consequences of the following investigations, a mechanism to explain the GO–AB cross-type nanostructure arrangement has been proposed, as shown in Fig. 11, which states that the GO nanosheets have their basal planes brightened with hydroxyl groups. These groups empower the proto analysis of the B–H bond in a portion of the AB molecules, which were present in the middle of the isolated GO sheets. This causes the development of the cationic AB connected to the oppositely charged oxygen through electrostatic interactions (Fig. 11B). On the other hand, the other AB atoms are not protonated and circulated inside the interlayer space, and the GO sheets are inclined to restack to make a sandwiched GO–AB–GO structure as the dissolvable material is evacuated. AB and its cationic initiator moved towards becoming encapsulated inside the GO interlayer.
For the current metal-catalyzed reaction, the initiation procedure happens on the metal catalyst surface, as proposed by the zero-order kinetic reaction. A plausible mechanism is shown in Fig. 12. It proposes that there is an interface between the AB molecules and the metal particle surface that creates an initiated complex species in the rate-determining step, on which attack by a water molecule readily results in the deliberate separation of the B–N bond and hydrolysis at the center of BH3 to form the borate ion together with H2 (eqn (1)). According to literature, without water, dehydrocoupling between the AB molecule occurs, which produces new B–N bonds, probably by means of a thorough connection in between, on the metal surface.46–50
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/d0ma00441c |
This journal is © The Royal Society of Chemistry 2020 |