Next Article in Journal
Spatiotemporal Evolution of Tourism Eco-Efficiency in Major Tourist Cities in China
Next Article in Special Issue
Pyramidal Solar Stills via Hollow Cylindrical Perforated Fins, Inclined Rectangular Perforated Fins, and Nanocomposites: An Experimental Investigation
Previous Article in Journal
Current Trends in Interprofessional Shared Decision-Making Programmes in Health Professions Education: A Scoping Review
Previous Article in Special Issue
Modified Nano-Fe2O3-Paraffin Wax for Efficient Photovoltaic/Thermal System in Severe Weather Conditions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Decomposition of Formic Acid and Acetic Acid into Hydrogen Using Graphitic Carbon Nitride Supported Single Metal Catalyst

by
Wan Nor Roslam Wan Isahak
1,*,
Muhammad Nizam Kamaruddin
1,
Zatil Amali Che Ramli
2,
Khairul Naim Ahmad
1,
Waleed Khalid Al-Azzawi
3 and
Ahmed Al-Amiery
1,4
1
Department of Chemical and Process Engineering, Faculty of Engineering and Built Environment, Universiti Kebangsaan Malaysia, Bangi 43600, Selangor, Malaysia
2
Fuel Cell Institute, Universiti Kebangsaan Malaysia, Bangi 43600, Selangor, Malaysia
3
Department of Medical Instruments Engineering Techniques, Al-Farahidi University, Baghdad 10001, Iraq
4
Energy and Renewable Energies Technology Center, University of Technology-Iraq, Baghdad 10001, Iraq
*
Author to whom correspondence should be addressed.
Sustainability 2022, 14(20), 13156; https://doi.org/10.3390/su142013156
Submission received: 29 August 2022 / Revised: 6 October 2022 / Accepted: 10 October 2022 / Published: 13 October 2022
(This article belongs to the Special Issue Advanced Technologies Applied to Renewable Energy)

Abstract

:
In a combination of generation and storage of hydrogen gas, both formic acid (FA) and acetic acid (AA) have been notified as efficient hydrogen carriers. This study was conducted to synthesize the monometallic catalysts namely palladium (Pd), copper (Cu), and zinc (Zn) on graphitic-carbon nitride (g-C3N4) and to study the potential of these catalysts in FA and mixed formic acid (FA)-acetic acid (AA) decomposition reaction. Several parameters have been studied in this work such as the type of active metals, temperature, and metal loadings. The mass percentage of Pd, Cu, and Zn metal used in this experiment are 1, 3, and 5 wt%, respectively. At low temperature of 30 °C, 5 wt% Pd/g-C3N4 catalyst yielded higher volume of gas with 3.3 mL, instead of other Pd percentage loadings. However, at higher temperature of 70 °C and 98% FA concentration, Pd with 1 wt%, 3 wt%, and 5 wt% of loading over g-C3N4 has successfully produced optimum gas (H2 and CO2) of 4.3 mL, 7.4 mL, and 4.5 mL in each reaction, respectively. At higher temperature, Pd metal showed high catalytic performance and the most active element of monometallic system in ambient condition. Meanwhile, at higher percentage of Pd metal, the catalytic decomposition reaction also increased thus producing more gas. However, it can be seen the agglomeration of the particles formed at higher loadings of Pd (5 wt%), and remarkably lowering the catalytic activity at higher temperature, while higher activity at low temperature of 30 °C. The result also showed low catalytic decomposition reaction for Cu and Zn catalyst, due to the small formation of Cu and Zn metal, but presence of high metal oxide (CuO) and (ZnO) promotes the passive layer formation on the catalyst surface.

1. Introduction

Global warming continues to gain attention from scientists around the world. Released of toxic and greenhouse gases such as sulphur dioxide (SO2) as well as carbon dioxide (CO2) are the main reason for global warming other than environmental pollution which caused by used of large fossil fuels. An important transformation has been made by scientists around the world to study alternative ways that are better and more effective in reducing the use of fossil fuels by using cleaner fuel sources such as hydrogen [1,2].
Hydrogen is one of the common elements and contributes 75% of the total mass of the universe. Hydrogen is not a source of energy or fuel, but it had been regarded as an energy carrier where it stores energy that had created elsewhere. For example, it reacts with oxygen in the air during combustion and produce heat that will be use as energy. There are various ways to produce hydrogen fuel such as electrolysis using hydroelectric, solar energy, wind energy, and nuclear energy. The essential benefits of hydrogen fuel used are carbon-free, non-toxic, lighter than air, and easily produced from a variety of different sources [3].
In a combination of generation and storage of hydrogen gas, both formic and acetic acid are known as efficient carrier fluid for hydrogen. Formic acid HCOOH is a type of carboxylic acid that contain 4.4 wt% hydrogen which is lower than DOE 2015 target up to 5.5 wt% [4]. Acetic acid is a type of molecule or a simple liquid with chemical formula of CH3COOH and it contains 6.7 wt% of hydrogen [5]. Formic acid and acetic commonly presence in organic waste from many processes especially biological process through fermentation to produce biogas or bio-hydrogen. High carboxylic acid content contributes to the environmental problem and utilization of these compounds can eliminate the related environmental issues in the future and turns it into clean hydrogen energy considered a significant value added to the process. Based on thermochemistry considerations, it is shown that formic acid can go through two different decomposition routes. However, decomposition into hydrogen and carbon dioxide gas would be more favorable at lower temperature, while at higher temperature dehydration of formic acid would be take place.
HCOOH → H2O + CO   ∆G298K = −14.9 kJ mol−1
HCOOH → H2 + CO2   ∆G298K = −35.0 kJ mol−1
Formic acid decomposition occurred at relatively slower rate. This kind of reaction needs a presence of catalysts to speed up the reaction by lowering the activation energy. It also makes the process of bonds breaking and formation of new bonds are more effective [1]. The active heterogeneous catalysts comprise of precious metals and transition metallic element such as Pd, Fe, Co, Ni, and Mn where the catalysts were used in decomposition of formic acid [6,7]. Metallic catalyst usually inhibited due to passive layer formation on the surface such as formate compound.
Graphitic carbon nitride (g-C3N4) is the most stable allotrope carbon nitride. It also has surface properties that can be used for various applications including as catalyst and photocatalyst reaction as it has high thermal and hydrothermal stability that enable it to function in a liquid, gas, and high temperature environment [8,9,10]. There are many literatures reported the synthesis method, potential, properties, advantageous, and catalytic/photocatalytic activities using graphitic carbon nitride (g-C3N4) material [11,12,13]. In photocatalytic reaction, graphitic carbon nitride has been widely used due to its properties such as suitable band structure, tunable bandgap, and low cost [14]. Nevertheless, modification must be done on bulk g-C3N4 to increase photocatalytic reaction rate. As described by Ji et al. (2020) in photocatalytic reaction, heterostructure with g-C3N4 can extend the adsorption to visible light region and retarded electron-hole pairs recombination [15]. While, introducing cyano groups to g-C3N4 nanosheets increase the charge carrier density, reduce the surface charge transfer resistance, and promote charge transfer and separation [16]. Qiu et al. (2022) reported presence of porous structure in g-C3N4 with many cyanide groups as a metal-free photocatalyst exhibited synergetic redox by photogenerated electron-hole pairs [17]. In addition, thermal calcination method using ammonium sulfate as porogen contributed to the creation of a porous structure in synthesized g-C3N4. Then, N-deficient ordered mesoporous g-C3N4 modified with AgPd nanoparticles with high catalytic and reusability was reported by a group of researcher, Wan et al., 2022 [18]. The high catalytic performance of this catalyst was recognized to the unique structure of N-ompg-C3N4 with higher surface area and abundant surface defects, the strong metal-support interaction between AgPd and N-ompg-C3N4, and charge transfer from Pd metal to Ag metal. Furthermore, g-C3N4 has also been applied in oxidative dehydrogenation of propane (ODHP). According to Chao et al. (2019), g-C3N4 can be a highly potential metal-free catalyst with high olefin selectivity in selective oxidative dehydrogenation of propane (ODHP) reaction [19]. Other than that, graphitic carbon nitride could be a potential support material for good dispersion of active metal catalysts. For example, addition of single atom Ru in Ru-g-C3N4 electrocatalyst showed high photocatalytic hydrogen production rate of 489.7 mmol gRu h−1, which attributed to the addition of single atom of Ru that able to alter electronic structure and improves HER performance, Yu et al., 2022 [20]. While Girma et al. 2021 demonstrated nanocomposite of graphitic carbon nitride and Cu/Fe as catalyst for CO2 reduction to CO with a maximum Faradaic efficiency of 84.4% at a low onset overpotential of −0.24 V vs. normal hydrogen electrode (NHE) [21]. In formic acid dehydrogenation, alloying of PdCo supported on g-C3N4 exhibited synergetic effect and enhanced the catalytic activity that achieved by the effect of alloying, small NPs size, and N-functionalities of the g-C3N4 support [22]. Homlamai et al. (2020) was also reported the utilization of non-noble metal can contribute to the formation of formate intermediate and needs higher temperature to reduce into CO2 and H2 [6]. Most of the catalytic decomposition of formic acid performed at fast rate and inhibited by formation of the formates species and it was irreversible.
In this study, heterogeneous catalysts especially single metallic catalysts will be synthesized and selected from noble metal and non-noble transition elements such as Pd, Cu, and Zn incorporated with graphitic carbon nitride (g-C3N4) as support materials. The synthesized catalysts will be characterized and further evaluated over the decomposition of short chain carboxylic acid such as formic acid, acetic acid, and mixture of them to form hydrogen gas. Performance study will be carried out using several parameters to optimize hydrogen production volume.

2. Materials and Methods

2.1. Materials and Reagents

Palladium (II) chloride (≥99.9 purity) precursor, zinc nitrate hexahydrate (reagent grade, 98%), copper (II) nitrate trihydrate (puriss. p.a., 99–104%), formic acid reagent (ACS reagent, ≥96%), melamine (99%) as graphitic carbon nitride source, sodium hydroxide (reagent grade, ≥98%,) pellets (anhydrous), sodium borohydride (Reagent Plus®, 99%), and ethanol (ACS reagent, ≥99.5%) was used during washing process. All chemicals were purchased from Sigma Aldrich Company (Darmstadt, Germany).

2.2. Synthesis of Modified Graphitic Carbon Nitride Based Monometallic Catalyst

To prepare 1 wt% Pd/g-C3N4 catalyst, a method by Wang et al. [23] was followed with some modifications. Moreover, 0.0433 g of PdCl2 and 1.98 g of g-C3N4 was added into 30 mL of distilled water. Then, the resulting suspensions was sonicated for 30 min and stirred for 4 h continuously using hot plate. Some amount of the 0.5M sodium hydroxide was added into the suspension to adjust the pH value at 8 to 9. Moreover, 172 mL of 1.0 mg/mL of sodium borohydride solution was poured into the suspension by droplets using a burette for reducing process where the metal precursor was reduced to metallic phase.
Then, the reduced suspension was re-sonicated for 20 min and the suspension was stirred for 1 h continuously by using a hot plate at 10 °C (using ice cubes) to ensure all the palladium chloride completely reduced to metallic phase of Pd. The resulting single metallic catalyst was filtered using centrifuge and washed with ethanol and distilled water in sequence to remove all the organic substance. Subsequently, the suspension was dried in the oven at 10 °C overnight. All the steps have been repeated to produce other catalysts with different amount of PdCl2 precursor according to the metal loadings of 3 wt% and 5 wt% of Pd over graphitic carbon nitride. The same method was applied to prepare the catalysts with 5 wt% metal loading, namely Cu/g-C3N4 and Zn/g-C3N4.

2.3. Catalysts Characterization

X-ray diffraction (XRD) diffractogram of the synthesized catalysts were recorded on a Bruker AXS D8 Advance (Karlsruhe, Germany) using a Cu Kα radiation source (40 kV, 40 mA). Scans were taken over the 2θ in range of 10° to 80° and wavelength, λ = 0.154 nm. Then, 1 g of sample was placed on the sample holder. The data obtain from the analysis was compared with standard peak data reported by Joint Committee on Powder Diffraction Standard (JCPDS) to match and identify the phase of the sample. The surface morphology and structural of the catalyst was observed by field emission scanning electron microscopy (FESEM) (Zeiss MERLIN, Oberkochen, Germany).

2.4. Catalysts Testing

Moreover, 0.03 g of Pd/g-C3N4 catalysts at different metal loadings were used in the reaction. Then, 1.0 mL of formic acid (98% concentration) was titrated into 5 mL 2-neck round bottle flask which connected with rubber tube as depicted in Figure 1. The experiment was observed for any potential reaction in 10 min to 1 h duration. For every 10 min, the gas trapped in the inverted measuring cylinder was measured based on displacement of water. These steps had repeated using different temperature of 50 °C and 70 °C. Subsequently, the steps were the same using different concentration of formic acid (70%, 50%, and 50%–50% formic acid-acetic acid mixture). The gas composition for the optimum reaction condition has been analyzed using gas chromatography (GC-TCD).
The last step is by repeating all the steps above using other catalysts, Cu/g-C3N4 and Zn/g-C3N4, according to the best suitable mass percentage chosen from (1 wt%, 3 wt%, and 5 wt%) Pd/g-C3N4. There are several parameters have been studied in this work such as type of active metal loadings (Pd/g-C3N4, Cu/g-C3N4, and Zn/g-C3N4), percentage of the metal loadings (1 wt%, 3 wt%, and 5 wt%), reactant concentration (98%, 70%, 50% formic acid, and 50%–50% of formic acid-acetic acid solution) and reaction temperature (30, 50, and 70 °C). Dilution of formic acid and acetic acid, TON, and TOF can be calculated using the equation below. The number of available Pd atoms was obtained with EDX analysis.
F o r m i c   a c i d   o r   a c e t i c   a c i d = D e s i r e d   d i l u t i o n   mL I n i t i a l   d i l u t i o n   mL × t o t a l   v o l u m e   o f   s o l u t i o n   mL
D i s t i l l e d   w a t e r   v o l u m e   = T o t a l   v o l u m e   o f   s o l u t i o n f o r m i c   o r   a c e t i c   a c i d   v o l u m e
TON   = M o l e   o f   r e a c t a n t × C o n v e r s i o n M o l e s   o f   P d   N P   n
TOF = TON T i m e   o f   r e a c t i o n   h

3. Results and Discussion

3.1. Catalysts Properties and Features

Five samples of catalysts had been synthesized using Pd, Zn, and Cu metals combined with graphitic carbon nitride which act as support. The Pd/g-C3N4 catalyst was synthesized in three mass percentage ratios (1 wt%, 3 wt%, and 5 wt% of metal in graphitic carbon nitride). For Cu/g-C3N4 and Zn/g-C3N4 catalysts, both were synthesized in a single mass percentage ratio (5 wt% of metal in graphitic carbon nitride). In addition, a control sample also included which consist of only g-C3N4.
For XRD analysis, all five samples of catalyst were analyzed. For all (1 wt%, 3 wt%, and 5 wt%) Pd/g-C3N4 catalyst, Figure 2 shows that all three Pd/g-C3N4 catalysts have the same highest XRD peak of graphitic carbon nitride (g-C3N4) which can detected between 25° and 30°. For palladium metal element (Pdo), all the three Pd/g-C3N4 catalysts have the same highest XRD peak which can be detected at 40°. There is also a small amount of alloy compound, palladium oxide (PdO) detected in the analysis. From the three samples, 5 wt% Pd/g-C3N4 catalyst has the highest XRD peak of Pdo compared to 1 wt% Pd/g-C3N4 and 3 wt% Pd/g-C3N4 catalyst.
For 5 wt% Cu/g-C3N4 catalyst, Figure 3 shows that highest XRD peak of graphitic carbon nitride (g-C3N4) can be detected between 25° and 30°. For copper metal element (Cu°), only few produced and no XRD peak of Cu metal. Instead, there were lot of alloy compound, copper oxide (CuO) can be seen in the analysis. The highest XRD peak for CuO can be detected between 35° and 40°.
For 5 wt% Zn/g-C3N4 catalyst, Figure 4 shows that highest XRD peak of graphitic carbon nitride (g-C3N4) can be detected between 25° and 30°. There are significant peaks represented to Zn2+ as ZnO have been detected without any metallic phase of Zn (Zn0). The highest XRD peak for ZnO can be detected between 35° and 40°.
Table 1 shows the percentage of crystallinity and amorphous as well as crystalline size of all synthesized catalyst. The average crystallite size of graphitic carbon nitride supported Pd catalyst ranges from 8 to 17 nm. Higher loadings of Pd gives higher crystallinity of 56.8%. The catalyst with 5 wt% of Zn has bigger average crystal size of 25 nm. It could contribute to the less dispersion of the particles towards less catalytic activity. It was comparable with the previous study that the average crystal size of activated carbon-based Pd catalyst was obtained within the range of 2 nm to 10 nm [23].
From physisorption study using BET method, it was found that higher loadings of Pd decrease the surface area of catalyst. Despite the low surface area of carbon nitride, it has several advantages as catalyst support. Its abundant tri-s-triazine groups can serve as anchoring points for metal precursors to produce small and well-dispersed NPs [22]. Apart from that, the rich electrons of nitrogen atoms in g-C3N4 support could transfer to Pd NPs and increase the electron cloud density of Pd to promote dissociation of hydrogen atom [24]. Surface properties analysis using BET method are summarized in Table 2.
For FESEM analysis, it only can be done for all three types of Pd/g-C3N4. Based on the EDX analysis, 1 wt% Pd/g-C3N4 has composition of 64.9 wt% nitrogen, 30.9 wt% carbon, 3.3 wt% oxygen, and 0.9 wt% palladium. Figure 5 shows the surface morphology of the catalyst at magnification of 10.00 KX. Based on the morphology analysis of all three catalyst of Pd/g-C3N4, the Pd particles are successfully dispersed (white spot) over the graphite-carbon nitride support with small crystal formation by using modification method from Wang et al. [23]. In FESEM images of each catalyst, graphitic carbon nitride in each catalyst samples also presence in layered structure, indicating the graphitic carbon nitride consisted of graphitic planes stacking along the c-axis, as also reported by previous literature [25].
Based on the EDX analysis, 3 wt% Pd/g-C3N4 has composition of 63.2 wt% nitrogen, 29.1 wt% carbon, 5.4 wt% oxygen, and 2.3 wt% palladium. Figure 6 shows the EDX analysis of the samples with different Pd loadings and have been summarized in Table 3.
Based on the EDX analysis, 5 wt% Pd/g-C3N4 has composition of 67.0 wt% nitrogen, 27.7 wt% carbon, 1.7 wt% oxygen, and 3.6 wt% palladium. Small amounts of oxygen content in the catalysts may represent the oxygenated phase of active metal as PdO. From the morphological analysis, Pd particles are widely dispersed over graphitic carbon nitride and all three catalysts have some small crystal formation on the surface of support material.

3.2. Catalytic Performance over Formic Acid, Acetic Acid and their Mixture Decomposition Reaction

Two parameters that are involved for the optimization in decomposition reaction are formic acid concentration, decomposition temperature and Pd loadings, mixture of formic acid-acetic acid, and effect of active metals, namely Pd, Cu, and Zn. For concentration of formic acid and mixture of formic acid-acetic acid, three types of catalysts were reacted with the formic acid at different concentration of 98%, 70%, and 50%, as well as a mixture of 50%–50% formic acid and acetic acid at constant temperature of 30 °C. Figure 7 shows 1 wt% Pd/g-C3N4 catalyst showed a reaction with all formic acid concentration. At higher concentration of 98% formic acid, the catalyst showed a better performance than lower concentrations. The active sites of the Pd surface were saturated with increasing formic acid concentrations, which might result in a decrease in hydrogen generation. Similar results were reported for formic acid decomposition to produce hydrogen by Mihet et al., 2020 [26] and Sanchez et al., 2018 [27].
Based on Figure 8, 1 wt% Pd/g-C3N4 catalyst was tested in 98% formic acid at 30 °C, 50 °C, and 70 °C. The catalyst performed well at 70 °C by producing gas (H2 and CO2), approximately of 4.3 mL which is much higher than the reaction at 30 °C and 50 °C. High temperature was increased the movement of the molecules and weaken the intramolecular bonds of formic acid. The presence of catalyst accelerating the C-H and o-H bonds breakout to form H2 and CO2.
Figure 9 shows 3 wt% Pd/g-C3N4 catalyst react with all types of concentration and at concentration 98% formic acid, and the catalyst shows better performance than other concentrations. Thus, 98% concentration of formic acid has been selected for further evaluation by other reaction parameters such as temperature and different Pd metal loadings as shown in Figure 10 and Figure 11.
Based on Figure 10, 3 wt% Pd/g-C3N4 catalyst was tested in 98% formic acid at 30, 50, and 70 °C. The catalyst performed well at 70 °C by producing gas (H2 and CO2), approximately 7.4 mL more than temperature at 30 and 50 °C. Figure 10 shows 5 wt% Pd/g-C3N4 catalyst react with all types of concentration and at concentration 98% formic acid, the catalyst shows better performance than other concentrations. Thus, 98% concentration of formic acid has been fixed for the second parameter which is temperature.
Based on Figure 12, 5 wt% Pd/g-C3N4 catalyst was tested in 98% formic acid at 30, 50, and 70 °C. The catalyst performed well at and 70 °C by producing gas (H2 and CO2), approximately 3.3 mL more than temperature at 30 and 50 °C.
Based on Figure 13, Figure 14 and Figure 15, it shows that 5 wt% Pd/g-C3N4 performed better compared to other catalyst. Moreover, 5 wt% Pd/g-C3N4 produced (H2 and CO2) much higher than other catalyst in all concentrations and temperatures. This may be due to the higher content of Pd metal in the 5 wt% Pd/g-C3N4 compared to other catalyst. It has been reported that Pd metal is the most active element or monometallic element in ambient condition and Pd metal also performed well in catalytic activity [28]. Thus, having high percentage of mass of Pd metal, the performance of decomposition reaction will increase. However, for higher temperature up to 70 °C, it showed a different trend while the total gas was decreased for 5 wt% Pd/C3N4 catalyst. It may be due to the early dehydration reaction of formic acid to produce PdCO or Pd formate caused the inhabitation of the catalyst [29].
Based on Figure 16a, it shows the conversion rate for 5 wt% Pd/g-C3N4 catalyst in decomposition reaction using 98% formic acid at different temperature (30 °C, 50 °C, and 70 °C) by assuming that 1 mol of formic acid produce 1 mol of gas (H2 and CO2) at STP. The highest conversion rate achieved by 5 wt% Pd/g-C3N4 catalyst is 0.741% at 70 °C. It showed significant value of formic acid conversion than other reported works from several researchers. Mihet et al. (2020) reported that templated carbon supported Pd catalyst did not showed any conversion of formic acid at room temperature [26]. The 5 wt% Pd/g-C3N4 catalyst was considered the best performance catalyst based on total gas production at room temperature. Since the higher temperature up to 70 °C did not show very significant in gas production, performance at room temperature might be more economical and practical. The optimum performance of 5 wt% Pd/g-C3N4 was shown at room temperature, compared with higher temperature of 50 and 70 °C. It indicated that the Pd atoms were stable at low temperature and less stable at higher temperature to produce possible interaction with formate and created a passive layer towards decreasing in the catalytic performance [24]. The measurement of kinetics was conducted over a range of temperature of 30 to 70 °C for 5 wt% Pd/g-C3N4. The Arrhenius plot in Figure 16b showed the activation energy was calculated to be 19.81 kJ/mol with the R2 of 0.986, indicating the kinetic equation was reasonable for the reaction.
In Table 4, a summarized results of catalysts’ performance from this study and other researchers is shown. In this study, the highest TOF−1 achieved is 1449.3 h−1, by 5 wt% Pd/g-C3N4 catalyst, while Cu and Zn based catalysts achieved significantly low TOF of 57.2 and 23.2 h−1, respectively. It was comparable with other findings [30,31] using Pd based heterogeneous catalysts. Highest Pd content of 5 wt% can provide more active site with most of the g-C3N4 support surface which has been covered with Pd species as shown in FESEM images (Figure 4). In heterogeneous catalysis, the highest TOF achieved is 7256 h−1, by palladium nanoparticles immobilized on carbon nanospheres developed by Zhu et al. [32]. High dispersion of fine Pd active metal contributed to the highest activity.
The decomposition reaction is continued for the control sample (g-C3N4) and other catalysts (5 wt% Cu/g-C3N4 and 5 wt% Zn/g-C3N4) by using 98% formic acid at temperature of 30 °C. Figure 17 shows the decomposition reaction of 98% formic acid with 5 wt% Pd/g-C3N4, 5 wt% Cu/g-C3N4, and 5wt% Zn/g-C3N4 at 30 °C. From Figure 16 below, 5 wt% Pd/g-C3N4 shows better performance compared to both of other catalyst. Moreover, 5 wt% Pd/g-C3N4 managed to produce gas about 3.3 mL compared to 5 wt% Cu/g-C3N4 and 5 wt% Zn/g-C3N4 as they produce very little gas. This may be due to the large formation of metal oxide formed based on the XRD analysis. The presence of metal oxide phase on the catalysts such as CuO and ZnO were recognized as less active species for the decomposition of carboxylic acid and could produce less gaseous products. The collected gas for 5 wt% metal loading catalysts have been analyzed using GC-TCD and the results are summarized in Table 5. Moreover, 5 wt% Pd/g-C3N4 showed highest hydrogen content up to 95.3% compared with Cu and Zn based catalysts. Lower hydrogen content due to higher CO2 production through formic acid decomposition. Meanwhile, no other gas such as carbon monoxide and methane has been detected.

4. Conclusions

Graphitic carbon nitride (g-C3N4) based materials have the potential to decompose short chain carboxylic acid into hydrogen gas. The effect of active metals such as Pd, Zn, and Cu gives an interesting finding that can be explored in the future. Reaction temperature showed a significant effect for formic acid-acetic acid decomposition. The optimum concentration and temperature for the catalyst to perform well are 98% concentration of formic acid and 70 °C, respectively. The g-C3N4 based catalysts with 5 wt% loading of the Pd metal is the most active catalyst compared with other catalysts with the highest gas production of 3.3, 4.4 and 4.5 mL at 30, 50, and 70 °C, respectively. Decomposition at room temperature gives additional advantages since no external heat sources are needed and it is much more practical to apply at any sites.

Author Contributions

Conceptualization, W.N.R.W.I.; methodology, M.N.K.; validation, W.N.R.W.I. and A.A.-A.; formal analysis, K.N.A. and Z.A.C.R.; investigation, M.N.K.; resources, W.N.R.W.I.; data curation, W.K.A.-A.; writing—original draft preparation, M.N.K. and W.N.R.W.I.; writing—review and editing, A.A.-A., Z.A.C.R., and W.N.R.W.I.; visualization, K.N.A.; supervision, W.N.R.W.I. and A.A.-A.; funding acquisition, W.N.R.W.I. All authors have read and agreed to the published version of the manuscript.

Funding

This research work was funded by the Ministry of Higher Education of Malaysia, and Universiti Kebangsaan Malaysia (UKM) under research code: FRGS/1/2020/TK0/UKM/02/31 and GUP-2020-012, respectively.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors wish to acknowledge Ministry of Higher Education for the financial support under research code: FRGS/1/2020/TK0/UKM/02/31, Universiti Kebangsaan Malaysia (UKM) for the financial support under research code: GUP-2020-012, and Centre for Research & Instrumentation Management (CRIM) for the support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ahmad, K.N.; Wan Isahak, W.N.R.; Rosli, M.I.; Yusop, M.R.; Kassim, M.B.; Yarmo, M.A. Rare earth metal doped nickel catalysts supported on exfoliated graphitic carbon nitride for highly selective CO and CO2 methanation. Appl. Surf. Sci. 2022, 571, 151321. [Google Scholar] [CrossRef]
  2. Azizi, M.A.H.; Dzakaria, N.; Isahak, W.N.R.W.; Yarmo, M.A. Effect of nickel on bimetallic nanoalloy catalyst for hydrogen generation. Malays. J. Anal. Sci. 2017, 21, 901–906. [Google Scholar] [CrossRef]
  3. Bahari, N.A.; Isahak, W.N.R.W.; Masdar, M.S.; Ba-Abbad, M.M. Optimization of the controllable crystal size of iron/zeolite nanocomposites using a Box–Behnken design and their catalytic activity. Appl. Nanosci. 2018, 9, 209–224. [Google Scholar] [CrossRef]
  4. Huang, Y.; Xu, J.; Ma, X.; Huang, Y.; Li, Q.; Qiu, H. An effective low Pd-loading catalyst for hydrogen generation from formic acid. Int. J. Hydrogen Energy 2017, 42, 18375–18382. [Google Scholar] [CrossRef]
  5. Cakiryilmaz, N.; Arbag, H.; Oktar, N.; Dogu, G.; Dogu, T. Catalytic performances of Ni and Cu impregnated MCM-41 and Zr-MCM-41 for hydrogen production through steam reforming of acetic acid. Catal. Today 2019, 323, 191–199. [Google Scholar] [CrossRef]
  6. Homlamai, K.; Maihom, T.; Choomwattana, S.; Sawangphruk, M.; Limtrakul, J. Single-atoms supported (Fe, Co, Ni, Cu) on graphitic carbon nitride for CO2 adsorption and hydrogenation to formic acid: First-principles insights. Appl. Surf. Sci. 2020, 499, 143928. [Google Scholar] [CrossRef]
  7. Wei, D.; Sang, R.; Sponholz, P.; Junge, H.; Beller, M. Reversible hydrogenation of carbon dioxide to formic acid using a Mn-pincer complex in the presence of lysine. Nat. Energy 2022, 7, 438–447. [Google Scholar] [CrossRef]
  8. Zhu, J.; Xiao, P.; Li, H.; Carabineiro, S.A.C. Graphitic carbon nitride: Synthesis, properties, and applications in catalysis. ACS Appl. Mater. Interfaces 2014, 6, 16449–16465. [Google Scholar] [CrossRef]
  9. Hasan, S.Z.; Ahmad, K.N.; Isahak, W.N.R.W.; Pudukudy, M.; Masdar, M.S.; Jahim, J.M. Synthesis, Characterisation and Catalytic Activity of NiO supported Al2O3 for CO2 Hydrogenation to Carboxylic Acids: Influence of Catalyst Structure. IOP Conf. Ser. Earth Environ. Sci. 2019, 268, 12079. [Google Scholar] [CrossRef]
  10. Ramli, Z.A.C.; Shaari, N.; Saharuddin, T.S.T. Progress and major BARRIERS of nanocatalyst development in direct methanol fuel cell: A review. Int. J. Hydrogen Energy 2022, 47, 22114–22146. [Google Scholar] [CrossRef]
  11. Harun, N.A.M.; Shaari, N.; Ramli, Z.A.C. Progress of g-C3N4 and carbon-based material composite in fuel cell application. Int. J. Energy Res. 2022, 12, 16281–16315. [Google Scholar] [CrossRef]
  12. Eswaran, M.; Dhanusuraman, R.; Tsai, P.C.; Ponnusamy, V.K. One-step preparation of graphitic carbon nitride/Polyaniline/Palladium nanoparticles based nanohybrid composite modified electrode for efficient methanol electro-oxidation. Fuel 2019, 251, 91–97. [Google Scholar] [CrossRef]
  13. Wang, S.; Li, Y.; Wang, X.; Zi, G.; Zhou, C.; Liu, B.; Liu, G.; Wang, L.; Huang, W. One-step supramolecular preorganization constructed crinkly graphitic carbon nitride nanosheets with enhanced photocatalytic activity. J. Mater. Sci. Technol. 2022, 104, 155–162. [Google Scholar] [CrossRef]
  14. Wang, X.; Liu, B.; Xiao, X.; Wang, S.; Huang, W. Boron dopant simultaneously achieving nanostructure control and electronic structure tuning of graphitic carbon nitride with enhanced photocatalytic activity. J. Mater. Chem. C 2021, 9, 14876–14884. [Google Scholar] [CrossRef]
  15. Ji, H.; Du, P.; Zhao, D.; Li, S.; Sun, F.; Duin, E.C.; Liu, W. 2D/1D graphitic carbon nitride/titanate nanotubes heterostructure for efficient photocatalysis of sulfamethazine under solar light: Catalytic “hot spots” at the rutile–anatase–titanate interfaces. Appl. Catal. B Environ. 2020, 263, 118357. [Google Scholar] [CrossRef]
  16. Wang, S.; Wang, X.; Liu, B.; Xiao, X.; Wang, L.; Huang, W. Boosting the photocatalytic hydrogen production performance of graphitic carbon nitride nanosheets by tailoring the cyano groups. J. Colloid Interface Sci. 2022, 610, 495–503. [Google Scholar] [CrossRef]
  17. Qiu, C.; Wang, S.; Zuo, J.; Zhang, B. Photocatalytic CO2 Reduction Coupled with Alcohol Oxidation over Porous Carbon Nitride. Catalysts 2022, 12, 672. [Google Scholar] [CrossRef]
  18. Wan, C.; Zhou, L.; Xu, S.; Jin, B.; Ge, X.; Qian, X.; Xu, L.; Chen, F.; Zhan, X.; Yang, Y.; et al. Defect engineered mesoporous graphitic carbon nitride modified with AgPd nanoparticles for enhanced photocatalytic hydrogen evolution from formic acid. Chem. Eng. J. 2022, 429, 132388. [Google Scholar] [CrossRef]
  19. Cao, L.; Dai, P.; Zhu, L.; Yan, L.; Chen, R.; Liu, D.; Gu, X.; Li, L.; Xue, Q.; Zhao, X. Graphitic carbon nitride catalyzes selective oxidative dehydrogenation of propane. Appl. Catal. B-Environ. 2020, 262, 118277. [Google Scholar] [CrossRef]
  20. Yu, Z.; Li, Y.; Torres-Pinto, A.; LaGrow, A.P.; Diaconescu, V.M.; Simonelli, L.; Sampaio, M.J.; Bondarchuk, O.; Amorim, I.; Araujo, A.; et al. Single-atom Ir and Ru anchored on graphitic carbon nitride for efficient and stable electrocatalytic/photocatalytic hydrogen evolution. Appl. Catal. B Environ. 2022, 310, 121318. [Google Scholar] [CrossRef]
  21. Woyessa, G.W.; dela Cruz, J.B.; Rameez, M.; Hung, C.-H. Nanocomposite catalyst of graphitic carbon nitride and Cu/Fe mixed metal oxide for electrochemical CO2 reduction to CO. Appl. Catal. B Environ. 2021, 291, 120052. [Google Scholar] [CrossRef]
  22. Navlani-García, M.; Salinas-Torres, D.; Mori, K.; Kuwahara, Y.; Yamashita, H. Enhanced formic acid dehydrogenation by the synergistic alloying effect of PdCo catalysts supported on graphitic carbon nitride. Int. J. Hydrogen Energy 2019, 44, 28483–28493. [Google Scholar] [CrossRef]
  23. Wang, X.; Tang, Y.; Gao, Y.; Lu, T. Carbon-supported Pd–Ir catalyst as anodic catalyst in direct formic acid fuel cell. J. Power Sources 2008, 175, 784–788. [Google Scholar] [CrossRef]
  24. Wang, Y.; Yao, J.; Li, H.; Su, D.; Antonietti, M. Highly Selective Hydrogenation of Phenol and Derivatives over a Pd@Carbon Nitride Catalyst in Aqueous Media. J. Am. Chem. Soc. 2011, 133, 2362–2365. [Google Scholar] [CrossRef]
  25. Li, Q.; He, Y.; Peng, R. Graphitic carbon nitride (g-C3N4) as a metal-free catalyst for thermal decomposition of ammonium perchlorate. RSC Adv. 2015, 5, 24507–24512. [Google Scholar] [CrossRef]
  26. Mihet, M.; Dan, M.; Barbu-Tudoran, L.; Lazar, M.D.; Blanita, G. Controllable H2 Generation by Formic Acid Decomposition on a Novel Pd/Templated Carbon Catalyst. Hydrogen 2020, 1, 22–37. [Google Scholar] [CrossRef]
  27. Sanchez, F.; Motta, D.; Roldan, A.; Hammond, C.; Villa, A.; Dimitratos, N. Hydrogen Generation from Additive-Free Formic Acid Decomposition Under Mild Conditions by Pd/C: Experimental and DFT Studies. Top. Catal. 2018, 61, 254–266. [Google Scholar] [CrossRef] [Green Version]
  28. Jones, S.; Kolpin, A.; Chi, S.; Tsang, E.; Jones, S.; Kolpin, A.; Chi, S.; Tsang, E. Modification of Pd for formic acid decomposition by support grafted functional groups Modification of Pd for formic acid decomposition by support grafted functional groups. Catal. Struct. React. 2015, 1, 19–24. [Google Scholar] [CrossRef] [Green Version]
  29. He, N.; Li, Z.H. Palladium-atom catalyzed formic acid decomposition and the switch of reaction mechanism with temperature. Phys. Chem. Chem. Phys. 2016, 18, 10005–10017. [Google Scholar] [CrossRef]
  30. Al-Nayili, A.; Majdi, H.S.; Albayati, T.M.; Saady, N.M.C. Formic Acid Dehydrogenation Using Noble-Metal Nanoheterogeneous Catalysts: Towards Sustainable Hydrogen-Based Energy. Catalysts 2022, 12, 324. [Google Scholar] [CrossRef]
  31. Caiti, M.; Padovan, D.; Hammond, C. Continuous Production of Hydrogen from Formic Acid Decomposition Over Heterogeneous Nanoparticle Catalysts: From Batch to Continuous Flow. ACS Catal. 2019, 10, 9188–9198. [Google Scholar] [CrossRef]
  32. Zhu, Q.-L.; Tsumori, N.; Xu, Q. Immobilizing Extremely Catalytically Active Palladium Nanoparticles to Carbon Nanospheres: A Weakly-Capping Growth Approach. J. Am. Chem. Soc. 2015, 137, 11743–11748. [Google Scholar] [CrossRef]
  33. Du, Y.; Shen, Y.-B.; Zhan, Y.-L.; Ning, F.-D.; Yan, L.-M.; Zhou, X.-C. Highly active iridium catalyst for hydrogen production from formic acid. Chin. Chem. Lett. 2017, 28, 1746–1750. [Google Scholar] [CrossRef]
Figure 1. Experimental set-up of carboxylic acid decomposition.
Figure 1. Experimental set-up of carboxylic acid decomposition.
Sustainability 14 13156 g001
Figure 2. XRD diffractogram of Pd/g-C3N4 catalyst at different metal loadings.
Figure 2. XRD diffractogram of Pd/g-C3N4 catalyst at different metal loadings.
Sustainability 14 13156 g002
Figure 3. XRD analysis of 5 wt% Cu/g-C3N4 catalyst.
Figure 3. XRD analysis of 5 wt% Cu/g-C3N4 catalyst.
Sustainability 14 13156 g003
Figure 4. XRD analysis of 5 wt% Zn/g-C3N4 catalyst.
Figure 4. XRD analysis of 5 wt% Zn/g-C3N4 catalyst.
Sustainability 14 13156 g004
Figure 5. Surface morphology of (a) 1 wt% Pd/g-C3N4, (b) 3 wt% Pd/g-C3N4, and (c) 5 wt% Pd/g-C3N4 at magnification of 10.00 KX.
Figure 5. Surface morphology of (a) 1 wt% Pd/g-C3N4, (b) 3 wt% Pd/g-C3N4, and (c) 5 wt% Pd/g-C3N4 at magnification of 10.00 KX.
Sustainability 14 13156 g005aSustainability 14 13156 g005b
Figure 6. EDX analysis of the catalyst samples, (a) 1 wt% Pd/ g-C3N4, (b) 3 wt% Pd/g-C3N4, and (c) 5 wt% Pd/g-C3N4.
Figure 6. EDX analysis of the catalyst samples, (a) 1 wt% Pd/ g-C3N4, (b) 3 wt% Pd/g-C3N4, and (c) 5 wt% Pd/g-C3N4.
Sustainability 14 13156 g006aSustainability 14 13156 g006b
Figure 7. Decomposition of formic acid (98%, 1 mL) and the mixture of formic-acetic acid (50%:50%, 1 mL) at different concentration using 1 wt% Pd/g-C3N4 at 30 °C.
Figure 7. Decomposition of formic acid (98%, 1 mL) and the mixture of formic-acetic acid (50%:50%, 1 mL) at different concentration using 1 wt% Pd/g-C3N4 at 30 °C.
Sustainability 14 13156 g007
Figure 8. Decomposition of 98% formic acid (1 mL) at different temperature over 1 wt% Pd/g-C3N4 catalyst.
Figure 8. Decomposition of 98% formic acid (1 mL) at different temperature over 1 wt% Pd/g-C3N4 catalyst.
Sustainability 14 13156 g008
Figure 9. Decomposition of formic acid (98%, 1 mL) and mixture of formic acid-acetic (50%:50%, 1 mL) at different concentration over 3 wt% Pd/g-C3N4 catalyst at 30 °C.
Figure 9. Decomposition of formic acid (98%, 1 mL) and mixture of formic acid-acetic (50%:50%, 1 mL) at different concentration over 3 wt% Pd/g-C3N4 catalyst at 30 °C.
Sustainability 14 13156 g009
Figure 10. Decomposition of 98% formic acid (1 mL) at different temperature over 3 wt% Pd/g-C3N4 catalyst.
Figure 10. Decomposition of 98% formic acid (1 mL) at different temperature over 3 wt% Pd/g-C3N4 catalyst.
Sustainability 14 13156 g010
Figure 11. Decomposition of formic acid (98%, 1 mL) and mixture of formic acid-acetic (50%:50%, 1 mL) at different concentration over 5 wt% Pd/g-C3N4 catalyst (0.03 g) at 30 °C.
Figure 11. Decomposition of formic acid (98%, 1 mL) and mixture of formic acid-acetic (50%:50%, 1 mL) at different concentration over 5 wt% Pd/g-C3N4 catalyst (0.03 g) at 30 °C.
Sustainability 14 13156 g011
Figure 12. Decomposition of 98% formic acid (1 mL) at different temperature using 5 wt% Pd/g-C3N4 catalyst.
Figure 12. Decomposition of 98% formic acid (1 mL) at different temperature using 5 wt% Pd/g-C3N4 catalyst.
Sustainability 14 13156 g012
Figure 13. Decomposition of formic acid (98%, 1 mL) and mixture of formic acid-acetic (50%:50%, 1 mL) at different concentration using (1 wt%, 3 wt%, and 5 wt%) Pd/g-C3N4 at 30 °C after 60 min.
Figure 13. Decomposition of formic acid (98%, 1 mL) and mixture of formic acid-acetic (50%:50%, 1 mL) at different concentration using (1 wt%, 3 wt%, and 5 wt%) Pd/g-C3N4 at 30 °C after 60 min.
Sustainability 14 13156 g013
Figure 14. Decomposition of formic acid (98%, 1 mL) and mixture of formic acid-acetic (50%:50%, 1 mL) at different concentration using (1 wt%, 3 wt%, and 5 wt%) Pd/g-C3N4 at 50 °C after 60 min.
Figure 14. Decomposition of formic acid (98%, 1 mL) and mixture of formic acid-acetic (50%:50%, 1 mL) at different concentration using (1 wt%, 3 wt%, and 5 wt%) Pd/g-C3N4 at 50 °C after 60 min.
Sustainability 14 13156 g014
Figure 15. Decomposition of formic acid (98%, 1 mL) and mixture of formic acid-acetic (50%:50%, 1 mL) at different concentration using (1 wt%, 3 wt%, and 5 wt%) Pd/g-C3N4 at 70 °C for 60 min.
Figure 15. Decomposition of formic acid (98%, 1 mL) and mixture of formic acid-acetic (50%:50%, 1 mL) at different concentration using (1 wt%, 3 wt%, and 5 wt%) Pd/g-C3N4 at 70 °C for 60 min.
Sustainability 14 13156 g015
Figure 16. (a) Conversion rate for 5 wt% Pd/g-C3N4 catalyst in decomposition reaction using 98% formic acid (1 mL) at different temperature (30 °C, 50 °C, and 70 °C) for 60 min. (b) Arrhenius plot of the reaction.
Figure 16. (a) Conversion rate for 5 wt% Pd/g-C3N4 catalyst in decomposition reaction using 98% formic acid (1 mL) at different temperature (30 °C, 50 °C, and 70 °C) for 60 min. (b) Arrhenius plot of the reaction.
Sustainability 14 13156 g016
Figure 17. Decomposition of 98% formic acid using 5 wt% Pd/g-C3N4, 5 wt% Cu/g-C3N4, and 5 wt% Zn/g-C3N4 at 30 °C.
Figure 17. Decomposition of 98% formic acid using 5 wt% Pd/g-C3N4, 5 wt% Cu/g-C3N4, and 5 wt% Zn/g-C3N4 at 30 °C.
Sustainability 14 13156 g017
Table 1. Percentage of crystallinity and amorphous as well as crystallite size of all synthesized catalyst.
Table 1. Percentage of crystallinity and amorphous as well as crystallite size of all synthesized catalyst.
Type of CatalystCrystallinity Percentage (%)Amorphous Percentage (%)Average Crystal Size of Metal/Metal Oxide (nm)
1 wt% Pd/g-C3N447.452.68.14
3 wt% Pd/g-C3N449.250.811.14
5 wt% Pd/g-C3N456.843.217.63
5 wt% Cu/g-C3N448.551.520.39
5 wt% Zn/g-C3N450.749.325.24
Table 2. Surface properties analysis using BET technique.
Table 2. Surface properties analysis using BET technique.
CatalystsBET Surface Area (m2/g)Micropore Area (m2/g)Pore Volume (cm3/g)Pore Size (nm)
GCN8.561.1890.01969.14
1% Pd/GCN7.551.760.01729.12
3% Pd/GCN7.321.620.01739.45
5% Pd/GCN7.181.350.01939.45
5% Cu/GCN10.271.770.02268.82
5% Zn/GCN10.251.540.02158.41
Table 3. Summarized of EDX analysis results of the samples at different Pd loadings.
Table 3. Summarized of EDX analysis results of the samples at different Pd loadings.
ElementsCatalysts with Different Pd Loadings (wt%)
1 wt% of Pd3 wt% of Pd5 wt% of Pd
Nitrogen64.963.267.0
Carbon30.929.127.7
Oxygen3.35.41.7
Palladium0.92.33.6
Table 4. Conversion, turnover frequency (TOF) and turnover number (TON) for 5 wt% Pd/g-C3N4, 5 wt% Cu/g-C3N4, and 5 wt% Zn/g-C3N4 catalyst (0.03 g) in decomposition reaction using 98% formic acid (1 mL) at 30 °C for 60 min.
Table 4. Conversion, turnover frequency (TOF) and turnover number (TON) for 5 wt% Pd/g-C3N4, 5 wt% Cu/g-C3N4, and 5 wt% Zn/g-C3N4 catalyst (0.03 g) in decomposition reaction using 98% formic acid (1 mL) at 30 °C for 60 min.
CatalystsFA Conversion (%)TONTOF (h−1)Total Gas Produced (mL) (1st Cycle)Total Gas Produced (mL) (2nd Cycle)Total Gas Produced (mL) (3rd Cycle)
5 wt% Pd/g-C3N40.5551449.31449.33.33.02.7
5 wt% Cu/g-C3N40.05157.257.20.3n.dn.d
5 wt% Zn/g-C3N40.02023.223.20.12n.dn.d
Pd/rGO [30]n.dn.d911.0
(10 mL of FA used)
80n.dn.d
Pd/C [31]9.5n.d1500 (at 50 °C, 5 bar)n.dn.dn.d
IrCp * Cl2bpym [33]n.dn.d2490 (at 40 °C, 300 mL of FA)350n.dn.d
n.d: Not determined; * denotes the presence of chiral carbon in the compound; Pd, Cu, and Zn atomic weight were estimated using EDX based on 0.03 g of total catalysts used.
Table 5. Gas composition analysis using GC-TCD for formic acid decomposition over different catalysts at 30 °C after 60 min.
Table 5. Gas composition analysis using GC-TCD for formic acid decomposition over different catalysts at 30 °C after 60 min.
CatalystsHydrogen (%)Carbon Dioxide (%)Other Gas (N2) (%)
5 wt% Pd/g-C3N495.34.7n.d
5 wt% Cu/g-C3N492.06.71.3
5 wt% Zn/g-C3N487.510.52.0
n.d: Not detected.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Isahak, W.N.R.W.; Kamaruddin, M.N.; Ramli, Z.A.C.; Ahmad, K.N.; Al-Azzawi, W.K.; Al-Amiery, A. Decomposition of Formic Acid and Acetic Acid into Hydrogen Using Graphitic Carbon Nitride Supported Single Metal Catalyst. Sustainability 2022, 14, 13156. https://doi.org/10.3390/su142013156

AMA Style

Isahak WNRW, Kamaruddin MN, Ramli ZAC, Ahmad KN, Al-Azzawi WK, Al-Amiery A. Decomposition of Formic Acid and Acetic Acid into Hydrogen Using Graphitic Carbon Nitride Supported Single Metal Catalyst. Sustainability. 2022; 14(20):13156. https://doi.org/10.3390/su142013156

Chicago/Turabian Style

Isahak, Wan Nor Roslam Wan, Muhammad Nizam Kamaruddin, Zatil Amali Che Ramli, Khairul Naim Ahmad, Waleed Khalid Al-Azzawi, and Ahmed Al-Amiery. 2022. "Decomposition of Formic Acid and Acetic Acid into Hydrogen Using Graphitic Carbon Nitride Supported Single Metal Catalyst" Sustainability 14, no. 20: 13156. https://doi.org/10.3390/su142013156

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