Next Article in Journal
Quantifying the Impact of Light on Ascorbic Acid Content in Lettuce: A Model Proposal
Previous Article in Journal
The Influence of Tourism’s Spatiotemporal Heterogeneity on the Urban–Rural Relationship: A Case Study of the Beijing–Tianjin–Hebei Urban Agglomeration, China
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effective Fuel Cell Electrocatalyst with Ultralow Pd Loading on Ni-N-Doped Graphene from Upcycled Water Bottle Waste

by
Aldona Balčiūnaitė
1,
Noha A. Elessawy
2,
Biljana Šljukić
3,
Arafat Toghan
4,5,
Sami A. Al-Hussain
4,
Marwa H. Gouda
6,
M. Elsayed Youssef
2 and
Diogo M. F. Santos
3,*
1
Department of Catalysis, Center for Physical Sciences and Technology, 10257 Vilnius, Lithuania
2
Computer Based Engineering Applications Department, Informatics Research Institute IRI, City of Scientific Research & Technological Applications (SRTA-City), Alexandria 21934, Egypt
3
Center of Physics and Engineering of Advanced Materials, Laboratory for Physics of Materials and Emerging Technologies, Chemical Engineering Department, Instituto Superior Técnico, Universidade de Lisboa, 1049-001 Lisbon, Portugal
4
Chemistry Department, College of Science, Imam Mohammad Ibn Saud Islamic University (IMSIU), Riyadh 11623, Saudi Arabia
5
Chemistry Department, Faculty of Science, South Valley University, Qena 83523, Egypt
6
Polymer Materials Research Department, Advanced Technology and New Materials Research Institute (ATNMRI), City of Scientific Research & Technological Applications (SRTA-City), Alexandria 21934, Egypt
*
Author to whom correspondence should be addressed.
Sustainability 2024, 16(17), 7469; https://doi.org/10.3390/su16177469
Submission received: 27 June 2024 / Revised: 23 August 2024 / Accepted: 26 August 2024 / Published: 29 August 2024
(This article belongs to the Section Energy Sustainability)

Abstract

:
Environmental pollution due to the excessive consumption of fossil fuels for energy production is a critical global issue. Fuel cells convert chemical energy directly into electricity in a clean and silent electrochemical process, but face challenges related to hydrogen storage, handling, and transportation. The direct borohydride fuel cell (DBFC), utilizing sodium borohydride as a liquid fuel, is a promising alternative to overcome such issues but requires the design of cost-effective nanostructured electrocatalysts. In this study, we synthesized nitrogen-doped graphene anchoring Ni nanoparticles (Ni@NG) by thermal degradation of polyethylene terephthalate bottle waste with urea and metallic Ni, and evaluated it as a sustainable carbon support. Electrocatalysts were prepared by incorporating ultralow amounts (0.09 to 0.27 wt.%) of Pd into the Ni@NG support. The resulting PdNi@NG electrocatalysts were characterized using ICP-OES, XPS, TEM, N2-sorption analysis, XRD, and Raman and FTIR spectroscopy. Voltammetry assessed the materials’ electrocatalytic activity for oxygen reduction and borohydride oxidation reactions in alkaline media, corresponding to the anodic and cathodic reactions in DBFCs. The electrocatalyst with 0.27 wt.% Pd loading (PdNi_15@NG) exhibited the best performance for both reactions. Consequently, it was employed as an anodic and cathodic material in a lab-scale DBFC, achieving a specific power of 3.46 kW gPd−1.

1. Introduction

Direct borohydride fuel cells (DBFCs) are a type of polymer electrolyte membrane (PEM) fuel cell that use sodium borohydride (NaBH4) as a fuel in alkaline media and hydrogen peroxide or oxygen as the oxidant. It was studied for the first time in 1962 by Indig and Snyder [1]. NaBH4 has gained a lot of attention from researchers due to its high specific capacity (5.7 Ahg−1) [2,3], low toxicity, safer transportation and storage, chemical stability, high hydrogen content (10.6 wt.%) [2,3], and high theoretical open circuit voltage (OCV) [3]. However, the commercialization of DBFCs is limited by using expensive noble-metal-based electrode materials as electrocatalysts [4,5,6,7,8]. One effective solution to the cost problem is developing highly active and stable anode and cathode electrocatalyst materials. Recent research has proven that borohydride oxidation reaction (BOR) on a Pd-based electrocatalyst has faster kinetics than on a Pt-based electrocatalyst. In addition to this, it has a lower onset potential than Au [9]. Pd-based electrocatalysts’ stability and electrocatalytic activity are improved by alloying them with lower-cost metals (e.g., Ni, Co, Fe, Sb, As, Bi, Ti) [10,11,12,13,14,15,16,17]. Among the nonprecious transition metals, nickel has unique qualities [18]. For example, the PdNi electrocatalyst’s active sites are renewed when OH-adsorbed groups are present on the nickel oxide surface. PdNi alloys’ enhanced activity and stability result from the synergistic interaction between Pd and Ni [9,14,15,19,20].
Hybrid materials with meso- or micropores have emerged as interrelated porous materials. Being connected to porous materials enhances structural porosity, specific surface area, mass transfer capacity, and shape compatibility [18]. In terms of different structural features [21], linked porous materials have unique homogenous morphologies. These substances could resemble sheets, as is the case with graphene, frequently used to support Pd-based electrocatalysts. Choosing a suitable electrocatalyst support material is critical to a bimetallic electrocatalyst’s stability, activity, and cost. Hosseini et al. prepared an anode composed of Pd/Co and Pd/Ni supported on nitrogen-doped reduced graphene oxide in a DBFC and realized power densities at 60 °C equal to 275.35 and 353.84 mW cm−2 [22].
Plastic waste is considered a dangerous environmental threat because it degrades slowly. Upcycling waste plastic is a cost-effective and simple technique to prepare electrocatalysts based on carbon [23,24,25,26]. Both the energy issue and insufficient management of plastic waste can be solved simultaneously by commercializing alternatives for clean energy supply devices and eliminating the accumulation of plastic waste. Graphene is commonly used as an electrode support in energy supply devices because of the high surface area of its layer structure, good conductivity, and sufficient thermal and chemical durability [27,28,29,30,31,32]. Generally, nitrogen-doped graphene metal-based electrocatalysts include a carbon backbone with defects containing a nitrogen/metal framework inserted into the graphitic/graphene-like structure [27]. When N atoms are incorporated into the graphene skeleton, the electron density in the nearby C atoms is rearranged, which changes the graphene geometry and improves its stability to donate electrons [33,34,35].
Here, nitrogen-doped graphene anchoring Ni nanoparticles (Ni@NG) is synthesized by a waste-to-value, scalable, one-pot pyrolysis of polyethylene terephthalate (PET) waste bottles with nickel metal and urea. This highly porous and stable substrate was used to support an ultralow amount of Pd nanoparticles, thus combining two approaches toward more efficient, cost-effective, and sustainable electrocatalysts. The electrocatalysts demonstrated improved performance as anodic and cathodic electrocatalysts in DBFCs, and therefore, fundamental studies were complemented with fuel cell testing, delivering high specific power densities. A literature search shows no previous work has been performed on creating high-performance NG from plastic wastes to support bimetallic PdNi nanocatalysts for DBFCs.

2. Materials and Methods

2.1. Synthesis of the Electrocatalysts

The samples were prepared by pyrolyzing waste PET drinking water bottles with urea and metallic nickel at 800 °C for 1 h. The produced black precipitate was collected and then ground to produce Ni@NG powder. PdNi@NG electrocatalysts with three different Pd metal nominal loading ratios, namely 0.5, 1, and 1.5 wt.%, were produced by mixing Ni@NG powder and tetrachloropalladinic acid solution (H2PdCl4.8wt.%H2O) purchased from Sigma-Aldrich (St. Louis, MO, USA). This was followed by adding 1 mL of a mixture of 1 M NaBH4 in 1 M NaOH solution and then stirring for 24 h. The obtained electrocatalyst was filtered, rinsed with deionized water until no chloride was detected, and then left to dry for 12 h at 60 °C. Figure 1 schematically shows the steps followed for preparing the PdNi@NG nanocomposites.

2.2. Characterization of the Electrocatalysts

Several methods were employed to characterize the prepared electrocatalysts, including XRD analysis (Schimadzu-7000, Columbia, MA, USA), Raman spectroscopy (SENTERRA, Bruker, Bremen, Germany), FTIR spectroscopy (Shimadzu FTIR-8400S, Kyoto, Japan), Brunauer–Emmett–Teller (BET) analysis, XPS analysis (K-Alpha, Thermo Fisher Scientific, Waltham, MA, USA), and transmission electron microscopy (TEM, JEOL JEM-1230, Tokyo, Japan). The metal loadings were determined using ICP-OES (Optima 7000DV, Perkin Elmer, Waltham, MA, USA).

2.3. Electrochemical Measurements

A PGSTAT100 potentiostat (Metrohm Autolab B.V., Utrecht, The Netherlands) was used to evaluate the electrocatalysts’ activity for BOR and ORR. Electrochemical measurements were performed in an electrochemical cell containing a glassy carbon electrode (GC) coated with the prepared electrocatalysts as the working electrode, a Pt sheet counter electrode, and a saturated calomel electrode (SCE) as the reference electrode. All recorded experiments were run in triplicate. The potentials given in the work refer to the SCE reference unless otherwise noted.
The electrocatalyst inks contained 5 mg of each synthesized electrocatalyst dispersed ultrasonically for 2 h in 100 μL of 2% solution of polyvinylidene fluoride (PVDF) in N–methyl–2–pyrrolidone (NMP). Then, 3 μL of the obtained ink was dropped on a glassy carbon electrode surface with a geometric area of 0.2 cm2 and dried at 80 °C for 4 h.
Cyclic voltammograms (CVs) were run in a 0.03 M NaBH4 + 2 M NaOH solution to assess the electrocatalysts’ activity for BOR. CVs were also recorded in deaerated (using an argon flow) and O2-saturated 0.1 M KOH to access their activity for ORR. Linear scan voltammograms (LSVs) were also recorded in O2-saturated 0.1 M KOH solution at a scan rate of 10 mV s−1 using the rotating disk electrode (RDE) method, with rotation rates ranging from 0 to 2400 rpm.
The activity of PdNi_15@NG as a DBFC anode and cathode electrocatalyst was examined at temperatures between 25 and 55 °C. The assembled DBFC used 1 M NaBH4 + 2 M NaOH solution as the fuel, O2-saturated 1 M NaOH as the oxidant solution, and a Fumasep FAA-3-PK-130 (FUMATECH BWT GmbH, Bietigheim-Bissingen, Germany) anion-exchange membrane (AEM) separating the two compartments (Figure 2). Each fuel cell compartment contained 100 mL of the corresponding aqueous electrolyte. The AEM membrane’s active area was ca. 30 cm2. All electrochemical measurements were performed with a Zennium electrochemical workstation (ZAHNER-Elektrik GmbH & Co. KG, Kronach, Germany). The performance of the fuel cell was evaluated by recording the cell polarization and obtaining the corresponding power density curves.

3. Results and Discussion

The present research work aimed to create sustainable and cost-effective nanostructured electrocatalysts for BOR and ORR in DBFCs. The masses of the elements (metal loadings) on the NG-based electrocatalyst were determined by ICP-OES analysis (Table 1). From ICP data, the fabricated bimetallic PdNi@NG electrocatalysts were determined to possess 0.09–0.27 wt.% Pd, which was lower than originally targeted. Still, these electrocatalysts with such low Pd amounts exhibited promising performance. In fact, high activity at low platinum-group metal loading is a general goal in the development of novel electrocatalysts, and thus, a further increase of Pd loading was not attempted.
Figure 3a displays the TEM image of PdNi_15@NG. Metallic particles are shown as dark spots scattered on irregularly, non-uniformly distributed graphitic sheets. Differences in the shape of metallic nanoparticles can be attributed to various factors, including heteroatomic intermix during crystal growth and differences in surface free energy. Pd and Ni nanoparticle sizes play an important role in the support interactions and the anchoring onto the N-doped graphene-based substrate. Although particle size estimation from the TEM image in Figure 3a indicates an average dimension at the nanoscale, it is not possible to quantify Pd and Ni particle sizes exactly.
As shown in Figure 3b, it is evident that the synthetic samples’ nitrogen adsorption-desorption isotherms have a distinctive IUPAC-type IV character. The Ni@NG adsorption isotherms exhibit meso- and microporosity, which is explained by both sheet defects and interlayer gaps between the graphene sheets. For PdNi_15@NG, the N2-sorption isotherm climbs above 0.9 relative pressure and does not reach saturation. It was also noticed that there is no meeting at the endpoint, referred to as “open hysteresis”. That phenomenon is well known for microstructure changes in carbonaceous materials, mainly when the N2 sorption occurs at low temperatures, entering the pores and leading to deformation. PdNi_5@NG and PdNi_10@NG also demonstrate the presence of meso- and macropores based on their significant specific volumes adsorbed at low relative pressures.
Table 2 displays the samples’ specific surface area, along with the total pore volume and the average diameter of the pores. The total pore volume is automatically reported, setting the final data point at the highest experimental relative pressure within the instrument’s range of accuracy. The average pore diameter in tenths of a nanometer was obtained using multilayer adsorption and capillary condensation of N2 gas within the narrow-diameter pores below a few tens of nanometers. The average pore diameter of PdNi_15@NG was found to be significantly higher than that of PdNi_10@NG and PdNi_5@NG.
The XPS elemental survey spectrum of the PdNi_15@NG surface is shown in Figure 4a. The graphitic C 1s peak was shown at 287.7 eV owing to the carbon bonds in the graphene architecture. Moreover, an N 1s peak at 400 eV demonstrates that the graphene was successfully doped with nitrogen. O, Ni, and Pd were also observed in the composite. As shown in Figure 4b, three peaks at 284.1, 285.3, and 287.4 eV may be identified in the high-resolution spectrum of C 1s. The last peak most likely originates from the C=O bond. That was confirmed with the high-resolution O 1s XPS spectrum (Figure 4c). The other two peaks correspond to sp3 hybridized single carbon bonds, C-C, and either C-O or C-N bonds, or a combination of both signals. As shown in Figure 4d, the high-resolution N 1s XPS spectrum provides further evidence of the presence of C-N bonds. The peak at 399.6 eV in the deconvoluted N 1s spectrum represents the pyridinic nitrogen groups [31,36]. Nitrogen-doped materials have been shown to have increased electrocatalytic activity toward BOR and ORR. Graphene has a larger density of π states near the Fermi level due to the pyridinic N, which can significantly increase ORR activity. The high-resolution Ni 2p XPS spectra shown in Figure 4e exhibit Ni with various chemical valences, detected at 855.6 eV, 861.5 eV, 873.1 eV, and 878.6 eV, corresponding to Ni2+ 2p3/2, Ni3+ 2p3/2, Ni2+ 2p1/2, and Ni3+ 2p1/2, respectively. These results confirm that the peak positions for the oxidation states of Ni2+ and Ni3+ are in close agreement with prior reports [37]. However, the strength of the Ni2+ 2p3/2 or 2p1/2 peaks is larger than that of the Ni3+ peaks. This indicates that there are more Ni2+ ions present than Ni3+ ions, which is beneficial for the electronic conductivity [3]. It is evident from the high-resolution spectra of Pd 3d shown in Figure 4f that Pd is mainly found as Pd0. As compared to the standard spectra of metal Pd0, the binding energies of Pd increased from 335.1 eV to 336.9 eV and from 340.4 eV to 341.1 eV corresponding to Pd 3d5/2 and Pd 3d3/2, respectively [38].
Raman spectroscopy confirmed the samples’ graphitic nature. Figure 5a shows that all the prepared electrocatalysts had two broad peaks corresponding to the G and D bands, at approximately 1590 cm−1 and 1370 cm−1, respectively. The sp3 hybridized carbon atoms, corresponding to the D band, are the reason for the structural defects. Conversely, organized sp2-linked carbon atom vibrations are the source of the G band, also known as the growth band [39]. The D band of PdNi_15@NG shows a shift to a higher wavelength contrasted with other PdNi@NG and Ni@NG electrocatalysts, and it could refer to the sustained damage caused to the C-C bond conjugation during the high-concentration Pd precipitation reaction [40]. However, several peaks were observed at low-frequency Raman bands, ascribed to laser-induced oxidation of the metal nanoparticles with a resultant formation of nickel oxide and palladium oxide. The degree of irregularity in the graphitic carbon structure, i.e., the concentration of the defect in the graphitic structure, is represented by the ID/IG intensity ratio. As can be seen from Figure 5a, the PdNi_15@NG sample shows the most disordered nanographitic structure with the ID/IG ratio of 0.93, i.e., higher than that of PdNi_10@NG (0.88), PdNi_5@NG (0.73), and Ni@NG (0.66), indicating the interaction between Ni@NG and Pd nanoparticles. This interaction is responsible for the increased degree of defects with an increase in the Pd concentration, and that causes this difference in peak ratio [40,41].
The processed samples’ FTIR spectra are displayed in Figure 5b, where a clear broad absorption at 3450 cm−1 is indicative of either the humidity absorbed during the samples’ preparation or the stretching vibration mode of –OH. The nitrogen-doped graphitic structure of the samples is evident by the presence of an in-plane C=C/C=N vibration at 1680 cm−1; this is a feature inherent in sp2 graphitic materials [42]. The C-O bond is attributed to the peak at roughly 664 cm−1. Moreover, two more peaks were seen between 610 and 714 cm−1; these peaks may be related to the stretching vibration of metal oxide/C–O–C [43].
Figure 5c shows the XRD patterns of the prepared electrocatalysts. The Ni face-centered cubic (fcc) structure was confirmed to be thermally incorporated into the graphitic structure by observations made at 2θ values of 44.2°, 51.5°, and 75.8°, corresponding to the (111), (200), and (220) planes, respectively (JCPDS No. 70-1849). PdNi_10@NG and PdNi_15@NG spectra show a distinctive Pd peak at 37.6° corresponding to the (111) plane (PDF 03-065-6174), and no discernible Pd diffraction peaks are seen in PdNi_5@NG; this could be caused by a low amount of Pd present that is not detectable by XRD [44].
Figure 6 shows the cyclic voltammograms of the prepared electrocatalysts recorded in the supporting 2 M NaOH electrolyte and in 0.03 M NaBH4 + 2 M NaOH solution, revealing their stable response with typical BH4 oxidation peaks. The peak current densities of the BH4 oxidation increase with the increase in Pd loading in the PdNi@NG electrocatalysts. The highest current density for the oxidation of BH4 ions was achieved using the PdNi_15@NG electrocatalyst (3.9 mAcm−2 at −0.24 V vs. SCE), being ~2.3 times higher than at PdNi_5@NG electrocatalyst.
With an increase in the potential scan rate from 5 to 1000 mVs−1, current density values increased from ~2 to 10 times for the prepared PdNi@NG electrocatalysts in the NaBH4 solution (Figure 7). The plots of the peak current (jp) versus the square root of scan rate ν1/2 were constructed and are shown in Figure 7d. As is evident, the peak current jp is linearly related to the square root of scan rate, ν1/2, with a good correlation coefficient, R2, of 0.907, 0.996, and 0.987 for PdNi_5@NG, PdNi_10@NG, and PdNi_15@NG, respectively. The shift of peak potential, Ep, to more positive values with the increasing scan rate is typical for irreversible processes. An increase of jp with ν1/2 and a slight shift of Ep suggest that the oxidation process of BH4 ions on the prepared electrocatalysts proceeds as an irreversible oxidation of a bulk species and is thus controlled by diffusion [45].
CVs were recorded in deaerated and O2-saturated 0.1 M KOH solutions (Figure 8), the latter showing the typical cathodic peak for ORR. Current densities obtained with PdNi_15@NG are ca. 1.5–2 times higher than for PdNi_10@NG, PdNi_5@NG, and Ni@NG electrocatalysts.
The RDE LSVs recorded in O2-saturated 0.1 M KOH (Figure 9) illustrate that the current densities increase with the rotation rate increase from 0 to 2000 rpm. With the rotation rate increase from 0 to 2000 rpm, the current density values increased ca. 2.7 to 4 times for the prepared PdNi@NG electrocatalysts in 0.1 M KOH solution. The obtained current densities are up to 2.1 times higher on PdNi_15@NG compared to the Ni@NG support and the other PdNi@NG electrocatalysts.
The number of exchanged electrons per O2 molecule, n, was calculated at different potentials in the 0.40–0.60 V range using the Koutecký–Levich (K–L) equation. The obtained K–L plots are provided in Figure 9a′–d′ and Figure 10b. A comparison of ORR LSVs recorded for the Ni@NG and different PdNi@NG electrocatalysts in O2-saturated 0.1 M KOH solution at 1600 rpm and K–L plots of the studied electrocatalysts determined at −0.5 V are presented in Figure 10a and Figure 10b, respectively. The studied samples are active ORR electrocatalysts in alkaline media, showing onset potential (Eonset, denoted as the potential at an ORR current density of −0.1 mA cm−2 [46]) values ranging from −0.213 to −0.320 V and corresponding half-wave potentials, E1/2, from −0.366 to −0.410 V (Figure 10a and Table 3).
The ORR RDE voltammograms at 10 mV s−1 and 1600 rpm were further used to obtain the Tafel plots (Figure 10c), as the Tafel slope is a key parameter in electrochemical kinetics that reflects the rate-determining step of a reaction. For the ORR, the Tafel slope provides insight into the reaction mechanism and the kinetics of the involved steps. Typically, lower Tafel slopes are associated with faster reaction kinetics and, thus, higher activity. However, in some cases, the highest activity may correlate with the highest Tafel slope for several potential reasons. Namely, the Tafel slope for Ni@NG was 0.100 V dec−1, lower than that for PdNi_5@NG (0.135 V dec−1), PdNi_10@NG (0.108 V dec−1), and PdNi_15@NG (0.157 V dec−1) (Table 3). While a high Tafel slope generally indicates slower reaction kinetics, the overall activity of a catalyst for the ORR can still be high due to factors such as surface properties, morphology, electronic effects, and mass transport considerations. Additionally, the electrocatalysts with higher Pd loading show more positive onset potentials, confirming their enhanced activity for ORR.
Namely, PdNi_15@NG electrocatalyst exhibited the highest ORR electrocatalytic activity, with Eonset of −0.161 V and E1/2 of −0.366 V; these values are 0.132 V and 0.044 V, respectively, shifted to more positive potentials than when using the Ni@NG electrocatalyst. The number of electrons transferred per O2 molecule (Figure 9d′) ranged between 3.2 and 3.6 in the 0.40–0.60 V potential range. This further suggests that ORR proceeds by mixed two- and four-electron mechanisms in the studied electrocatalysts.
The best performance of the PdNi_15@NG electrocatalyst for both BOR and ORR results from the highest Pd loading and the highest number of defects in its structure, accelerating the charge transfer process. Moreover, this electrocatalyst’s significantly higher pore volume and pore size diameter enable the unobstructed flow of reactants (BH4, O2) and products (BO2, OH) to and from the Pd and Ni active centers, enhancing the mass transfer process.
The practical applicability of PdNi_15@NG as a fuel cell electrocatalyst was examined in an alkaline single-cell DBFC using NaBH4 as the fuel and O2 as the oxidant. This DBFC was constructed using PdNi_15@NG both as an anode and cathode electrocatalyst. Figure 11 presents the fuel cell polarization and corresponding power density curves at 25, 35, 45, and 55 °C. All polarization curves were run in triplicate without any visible loss of activity; curves overlapped for each repetition, showing that the PdNi_15@NG electrocatalyst exhibited good stability.
The main determined fuel cell parameters are summarized in Table 4. The lab DBFC exhibited an open circuit voltage close to 1.0 V. At 25 °C, the DBFC attained a peak power density of 7.1 mW cm−2 at a current density of 14 mA cm−2 and a cell voltage of 0.50 V. Moreover, the peak power density values increased by ca. 1.3 times when increasing the temperature from 25 to 55 °C. Considering the ICP-OES data (cf. Table 1), the Pd loading on the prepared PdNi_15@NG-based fuel cell electrodes was 2.05 µgPd cm−2. This leads to a specific power of 3.46 kW gPd−1, significantly higher than for other systems reported in the literature using Pd-based anode electrocatalysts (Table 5).

4. Conclusions

The present study used one-pot pyrolysis of PET waste bottles with nickel metal and urea as a fast and simple process of preparing nickel on N-doped graphene, used as a sustainable support material for anchoring ultralow loadings of palladium nanoparticles. This straightforward method has substantial benefits, including simple operating stages and reaction setup, allowing scaling up to an industrial scale.
From the produced PdNi@NG electrocatalysts, PdNi_15@NG exhibited the highest electrocatalytic activity for borohydride oxidation and oxygen reduction reactions. This electrocatalyst achieved a current density of 3.9 mA cm−2 at −0.24 V vs. SCE for the oxidation of BH4 in 0.03 M NaBH4 + 2 M NaOH solution. As for ORR in alkaline media, it exhibited an Eonset value of −0.213 V vs. SCE and an E1/2 of −0.366 V vs. SCE, with a number of electrons transferred of 3.2–3.6 being determined by K-L analysis.
The direct borohydride fuel cells (DBFCs) assembled using PdNi_15@NG as an anode and cathode electrocatalyst in 1 M NaBH4 + 2 M NaOH anolyte and O2-saturated 1 M NaOH catholyte attained a peak power density of 7.1 mW cm−2 at a current density of 14 mA cm−2 and a cell voltage of 0.50 V at 25 °C. Based on the Pd loading, this corresponds to an outstanding specific power of 3.46 kW gPd−1. Accordingly, these data confirm the prepared PdNi@NG electrocatalysts, especially PdNi_15@NG, exhibit promising performance as potential anode and cathode electrode materials for DBFCs. Moreover, N-doped graphene derived from upcycled PET bottle waste was established as a cost-effective and sustainable electrocatalyst support.

Author Contributions

Conceptualization, N.A.E. and D.M.F.S.; methodology, N.A.E., A.B., B.Š. and D.M.F.S.; validation, A.T., S.A.A.-H., B.Š. and D.M.F.S.; formal analysis, A.B., A.T. and S.A.A.-H.; investigation, A.B., N.A.E. and M.H.G.; writing—original draft preparation, A.B., N.A.E. and D.M.F.S.; writing—review and editing, M.E.Y., B.Š. and D.M.F.S.; visualization, M.E.Y., A.T. and D.M.F.S.; supervision, D.M.F.S.; project administration, N.A.E., M.H.G., M.E.Y. and D.M.F.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Fundação para a Ciência e a Tecnologia (FCT, Portugal) through project EXPL/EQU-EQU/0517/2021. FCT also funded a research contract in the scope of pro-grammatic funding UIDP/04540/2020 (D.M.F. Santos) and contract no. IST-ID/156-2018 (B. Šljukić). Thanks to the Academy of Scientific Research and Technology and Bibliotheca Alexandria (ASRT-BA) for supporting this research through post-doctoral research grant project No. 1436.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in the study are included in the article; further inquiries can be directed to the corresponding author.

Acknowledgments

This publication is based upon work from COST Action Waste biorefinery technologies for accelerating sustainable energy processes (WIRE), CA20127, supported by COST (European Cooperation in Science and Technology); www.cost.eu.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Indig, M.E.; Snyder, R.N. Sodium Borohydride, An Interesting Anodic Fuel (1). J. Electrochem. Soc. 1962, 109, 1104–1106. [Google Scholar] [CrossRef]
  2. Qin, H.; Jiang, L.; He, Y.; Liu, J.; Cao, K.; Wang, J.; He, Y.; Ni, H.; Chi, H.; Ji, Z. Carbon Supported Silver Nanowires with Enhanced Catalytic Activity and Stability Used as a Cathode in a Direct Borohydride Fuel Cell. J. Mater. Chem. A 2013, 1, 15323–15328. [Google Scholar] [CrossRef]
  3. Chatenet, M.; Micoud, F.; Roche, I.; Chainet, E. Kinetics of Sodium Borohydride Direct Oxidation and Oxygen Reduction in Sodium Hydroxide Electrolyte: Part II. O2 Reduction. Electrochim. Acta 2006, 51, 5459–5467. [Google Scholar] [CrossRef]
  4. Ma, J.; Sahai, Y.; Buchheit, R.G. Direct Borohydride Fuel Cell Using Ni-Based Composite Anodes. J. Power Sources 2010, 195, 4709–4713. [Google Scholar] [CrossRef]
  5. Guo, Y.; Hu, Z.; Cao, Y.; Tan, Q.; Yang, D.; Che, Y.; Zhang, C.; Ming, P.; Xiao, Q. Boosting borohydride oxidation kinetics by manipulating hydrogen evolution and oxidation through octahedral Pt–Ni/C for high-performance direct borohydride fuel cells. J. Power Sources 2024, 612, 234786. [Google Scholar] [CrossRef]
  6. Zhiani, M.; Mohammadi, I. Performance Study of Passive and Active Direct Borohydride Fuel Cell Employing a Commercial Pd Decorated Ni-Co/C Anode Catalyst. Fuel 2016, 166, 517–525. [Google Scholar] [CrossRef]
  7. Milikic, J.; Martins, M.; Dobrota, A.S.; Bozkurt, G.; Soylu, G.S.P.; Yurtcan, A.B.; Skorodumova, N.V.; Pasti, I.A.; Sljukic, B.; Santos, D.M.F. A Pt/MnV2O6 Nanocomposite for the Borohydride Oxidation Reaction. J. Energy Chem. 2021, 55, 428–436. [Google Scholar] [CrossRef]
  8. Wang, B.; Zhang, D.; Ye, K.; Cheng, K.; Cao, D.; Wang, G.; Cheng, X. Plastic Supported Platinum Modified Nickel Electrode and its High Electrocatalytic Activity for Sodium Borohydride Electrooxidation. J. Energy Chem. 2015, 24, 497–502. [Google Scholar] [CrossRef]
  9. Ma, J.; Sahai, Y. Direct Borohydride Fuel Cells-Current Status, Issues, and Future Directions. In Anion Exchange Membrane Fuel Cells: Principles, Materials and Systems; An, L., Zhao, T.S., Eds.; Springer: Cham, Germany, 2018; pp. 249–283. [Google Scholar]
  10. Martins, M.; Šljukić, B.; Metin, O.; Sevim, M.; Sequeira, C.A.C.; Şener, T.; Santos, D.M.F. Bimetallic PdM (M = Fe, Ag, Au) Alloy Nanoparticles Assembled on Reduced Graphene Oxide as Catalysts for Direct Borohydride Fuel Cells. J. Alloys Compd. 2017, 718, 204–214. [Google Scholar] [CrossRef]
  11. Hosseini, M.G.; Mahmoodi, R. Improvement of Energy Conversion Efficiency and Power Generation in Direct Borohydride Hydrogen Peroxide Fuel Cell: The Effect of Ni-M Core-Shell Nanoparticles (M = Pt, Pd, Ru)/Multiwalled Carbon Nanotubes on the Cell Performance. J. Power Sources 2017, 370, 87–97. [Google Scholar] [CrossRef]
  12. Behmenyar, G.; Akın, A.N. Investigation of Carbon Supported Pd-Cu Nanoparticles as Anode Catalysts for Direct Borohydride Fuel Cell. J. Power Sources 2014, 249, 239–246. [Google Scholar] [CrossRef]
  13. Duan, D.; You, X.; Liang, J.; Liu, S.; Wang, Y. Carbon Supported Cu-Pd Nanoparticles as Anode Catalyst for Direct Borohydride-Hydrogen Peroxide Fuel Cells. Electrochim. Acta 2015, 176, 1126–1135. [Google Scholar] [CrossRef]
  14. Calderón, J.C.; Rios Ráfales, M.; Nieto-Monge, M.J.; Pardo, J.I.; Moliner, R.; Lázaro, M.J. Oxidation of CO and Methanol on Pd-Ni Catalysts Supported on Different Chemically-Treated Carbon Nanofibers. Nanomaterials 2016, 6, 187. [Google Scholar] [CrossRef] [PubMed]
  15. Hosseini, M.G.; Mahmoodi, R. Ni@M (M = Pt, Pd and Ru) Core@Shell Nanoparticles on a Vulcan XC-72R Support with Superior Catalytic Activity toward Borohydride Oxidation: Electrochemical and Fuel Cell Studies. New J. Chem. 2017, 41, 13408–13417. [Google Scholar] [CrossRef]
  16. He, P.; Wang, X.; Fu, P.; Wang, H.; Yi, L. The Studies of Performance of the Au Electrode Modified by Zn as the Anode Electrocatalyst of Direct Borohydride Fuel Cell. Int. J. Hydrogen Energy 2011, 36, 8857–8863. [Google Scholar] [CrossRef]
  17. Yang, Q.; Zhang, L.; Ma, S.; Wang, W.; Zhao, M.; Yuan, W.; Li, C. Self-assembled TiO2 Nanopore-Confined Growth of Pd NPs on Pristine Graphene for Superior Electrocatalytic Performance toward Formic Acid Oxidation. ACS Appl. Energy Mater. 2023, 6, 9318–9327. [Google Scholar] [CrossRef]
  18. Ahmad, A.; Noor, A.; Anwar, A.; Majeed, S.; Khan, S.; Nisa, Z.; Ali, S.; Gnanasekaran, L.; Rajendran, S.; Li, H. Support based metal incorporated layered nanomaterials for photocatalytic degradation of organic pollutants. Environ. Res. 2024, 260, 119481. [Google Scholar] [CrossRef] [PubMed]
  19. Mahmoodi, R.; Chalani, M.; Hosseini, M.G.; Darbandi, M. Novel Electrocatalysts for Borohydride Fuel Cells: Enhanced Power Generation by Optimizing Anodic Core-Shell Nanoparticles on Reduced Graphene Oxide. New J. Chem. 2020, 44, 11974–11987. [Google Scholar] [CrossRef]
  20. Mahmoodi, R.; Hosseini, M.G.; Rasouli, H. Enhancement of Output Power Density and Performance of Direct BorohydrideHydrogen Peroxide Fuel Cell Using Ni-Pd Core-Shell Nanoparticles on Polymeric Composite Supports (rGO-PANI) as Novel Electrocatalysts. Appl. Catal. B 2019, 251, 37–48. [Google Scholar] [CrossRef]
  21. Fan, X.; Yuan, W.; Zhang, D.; Li, C. Heteropolyacid-Mediated Self-Assembly of Heteropolyacid-Modified Pristine Graphene Supported Pd Nanoflowers for Superior Catalytic Performance toward Formic Acid Oxidation. ACS Appl. Energy Mater. 2018, 1, 411–420. [Google Scholar] [CrossRef]
  22. Hosseini, M.G.; Esfahlan, V.D.; Wolf, S.; Hacker, V. Novel Bimetallic Pd–X (X = Ni, Co) Nanoparticles Assembled on N-Doped Reduced Graphene Oxide as an Anode Catalyst for Highly Efficient Direct Sodium Borohydride–Hydrogen Peroxide Fuel Cells. ACS Appl. Energy Mater. 2021, 4, 6025–6039. [Google Scholar] [CrossRef]
  23. Park, J.C.; Kim, J.C.; Park, S.; Kim, D.W. Efficient waste polyvinyl(butyral) and cellulose composite enabled carbon nanofibers for oxygen reduction reaction and water remediation. Appl. Surf. Sci. 2020, 510, 145505. [Google Scholar] [CrossRef]
  24. Daniel, G.; Kosmala, T.; Dalconi, M.C.; Nodari, L.; Badocco, D.; Pastore, P.; Lorenzetti, A.; Granozzi, G.; Durante, C. Upcycling of polyurethane into iron-nitrogen-carbon electrocatalysts active for oxygen reduction reaction. Electrochim. Acta 2020, 362, 137200. [Google Scholar] [CrossRef]
  25. Elessawy, N.A.; Elnouby, M.; Gouda, M.; Mohy Eldin, M.S.; Farag, H.A.; Konsowa, A.H. Simple Self-assembly Synthesis for Cost-Effective Alkaline Fuel Cell Bi-functional Electrocatalyst Synthesized from Polyethylene Terephthalate Waste Bottles. J. Electron. Mater. 2020, 49, 1009–1016. [Google Scholar] [CrossRef]
  26. Bhaskaran, R.; Chetty, R. One-Pot Room Temperature Synthesis of Nitrogen-Doped Graphene and Its Application as Catalyst Support for ORR in PEMFCs. ACS Appl. Energy Mater. 2024, 7, 390–402. [Google Scholar] [CrossRef]
  27. Muhyuddin, M.; Mustarelli, P.; Santoro, C. Recent Advances in Waste Plastic Transformation into Valuable Platinum-Group Metal-Free Electrocatalysts for Oxygen Reduction Reaction. ChemSusChem 2021, 14, 3785. [Google Scholar] [CrossRef]
  28. He, D.; Tang, H.; Kou, Z.; Pan, M.; Sun, X.; Zhang, J.; Mu, S. Engineered graphene materials: Synthesis and applications for polymer electrolyte membrane fuel cells. Adv. Mater. 2017, 29, 1601741. [Google Scholar] [CrossRef]
  29. Fang, Y.; Hu, R.; Liu, B.; Zhang, Y.; Zhu, K.; Yan, J.; Ye, K.; Cheng, K.; Wang, G.; Cao, D. MXene-derived TiO2/reduced graphene oxide composite with an enhanced capacitive capacity for Li-ion and K-ion batteries. J. Mater. Chem. A 2019, 7, 5363–5372. [Google Scholar] [CrossRef]
  30. Kou, Z.; Meng, T.; Guo, B.; Amiinu, I.S.; Li, W.; Zhang, J.; Mu, S. A generic conversion strategy: From 2D metal carbides (MxCy) to M-self-doped graphene toward high efficiency energy applications. Adv. Funct. Mater. 2017, 27, 1604904. [Google Scholar] [CrossRef]
  31. Gouda, M.H.; Elnouby, M.; Aziz, A.N.; Youssef, M.E.; Santos, D.M.F.; Elessawy, N.A. Green and low-cost membrane electrode assembly for proton exchange membrane fuel cells: Effect of double-layer electrodes and gas diffusion layer. Front. Mater. 2020, 6, 337. [Google Scholar] [CrossRef]
  32. Elessawy, N.A.; Backović, G.; Janesuda, H.; Martins, M.; Rakočević, L.; Gouda, M.H.; Toghan, A.; Youssef, M.E.; Šljukić, B.; Santos, D.M.F. From PET Bottles Waste to N-Doped Graphene as Sustainable Electrocatalyst Support for Direct Liquid Fuel Cells. Catalysts 2023, 13, 525. [Google Scholar] [CrossRef]
  33. Wu, G.; Santandreu, A.; Kellogg, W.; Gupta, S.; Ogoke, O.; Zhang, H.; Wang, H.L.; Dai, L. Carbon nanocomposite catalysts for oxygen reduction and evolution reactions: From nitrogen doping to transition-metal addition. Nano Energy 2016, 29, 83–110. [Google Scholar] [CrossRef]
  34. Mohamed, M.; Carrasco-Marín, F.; Elessawy, N.A.; Hamad, H.A.F. Glucose Derived N-Doped Graphitic Carbon: Facile One-Pot Graphitic Structure-Controlled Chemical Synthesis with Comprehensive Insight into the Controlling Mechanisms. ChemistrySelect 2020, 5, 14685. [Google Scholar] [CrossRef]
  35. Saini, R.; Naaz, F.; Bashal, A.H.; Pandit, A.H.; Farooq, U. Recent advances in nitrogen-doped graphene-based heterostructures and composites: Mechanism and active sites for electrochemical ORR and HER. Green Chem. 2024, 26, 57–102. [Google Scholar] [CrossRef]
  36. Osman, A.I.; Farrell, C.; Al-Muhtaseb, A.H.; Al-Fatesh, A.S.; Harrison, J.; Rooney, D.W. Pyrolysis kinetic modelling of abundant plastic waste (PET) and in-situ emission monitoring. Environ. Sci. Eur. 2020, 32, 112. [Google Scholar] [CrossRef]
  37. Dubey, P.; Kaurav, N.; Devan, R.S.; Okram, G.S.; Kuo, Y.K. The effect of stoichiometry on the structural, thermal and electronic properties of thermally decomposed nickel oxide. RSC Adv. 2018, 8, 5882. [Google Scholar] [CrossRef] [PubMed]
  38. Yang, S.; Dong, J.; Yao, Z.; Shen, C.; Shi, X.; Tian, Y.; Lin, S.; Zhang, X. One-Pot Synthesis of Graphene-Supported Monodisperse Pd Nanoparticles as Catalyst for Formic Acid Electro-oxidation. Sci. Rep. 2014, 4, 4501. [Google Scholar] [CrossRef]
  39. Katubi, K.M.M.; Alsaiari, N.S.; Alzahrani, F.M.; Siddeeg, S.A.; Tahoon, M. Synthesis of Manganese Ferrite/Graphene Oxide Magnetic Nanocomposite for Pollutants Removal from Water. Processes 2021, 9, 589. [Google Scholar] [CrossRef]
  40. Ji, J.; Dong, P.; Lin, Y.; Zeng, X.; Li, X.; Yang, X.; He, Q.; Zhang, Y.; Xu, M. One-pot synthesis of PdM/RGO (M=Co, Ni, or Cu) catalysts under the existence of PEG for electro-oxidation of methanol. J. Nanopart. Res. 2018, 20, 192. [Google Scholar] [CrossRef]
  41. Minh Dat, N.; Linh, V.N.P.; Huy, L.A.; Huong, N.T.; Tu, T.H.; Phuong, N.T.L.; Nam, H.M.; Thanh Phong, M.; Hieu, N.H. Fabrication and antibacterial activity against Pseudomonas aeruginosa and Staphylococcus aureus of silver nanoparticle decorated reduced graphene oxide nanocomposites. Mater. Technol. 2019, 34, 36. [Google Scholar] [CrossRef]
  42. Mohamed, M.; Elessawy, N.; Carrasco-Marínc, F.; Hamad, H. A novel one-pot facile economic approach for the mass synthesis of exfoliated multilayered nitrogen-doped graphene-like nanosheets: New insights into the mechanistic study. Phys. Chem. Chem. Phys. 2019, 21, 13611. [Google Scholar] [CrossRef] [PubMed]
  43. Abdul Mannan, M.; Hirano, Y.; Quitain, A.T.; Koinuma, M.; Kida, T. Graphene Oxide to B, N Co-doped Graphene through Tris-dimethylaminoborane Complex by Hydrothermal Implantation. Am. J. Mater. Sci. 2019, 9, 22–28. [Google Scholar]
  44. Elessawy, N.A.; Toghan, A.; Elnouby, M.S.; Alakhras, A.I.; Hamad, H.A.; Youssef, M.E. Development and activity enhancement of zirconium/vanadium oxides as micro-heterogeneous ceramic electrocatalyst for ORR in low-temperature fuel cell. Ceram. Int. 2023, 49, 4313–4321. [Google Scholar] [CrossRef]
  45. Bard, A.J.; Faulkner, L.R. Electrochemical Methods Fundamentals and Applications, 2nd ed.; Wiley: New York, NY, USA, 2004. [Google Scholar]
  46. Monini, V.; Bonechi, M.; Bazzicalupi, C.; Bianchi, A.; Gentilesca, P.; Giurlani, W.; Innocenti, M.; Meoli, A.; Romano, G.M.; Savastano, M. Oxygen reduction reaction (ORR) in alkaline solution catalysed by an atomically precise catalyst based on a Pd(ii) complex supported on multi-walled carbon nanotubes (MWCNTs). Electrochemical and structural considerations. Dalton Trans. 2024, 53, 2487–2500. [Google Scholar] [CrossRef]
  47. Montserrat-Sisó, G.; Wickman, B. PdNi thin films for hydrogen oxidation reaction and oxygen reduction reaction in alkaline media. Electrochim. Acta 2022, 420, 140425. [Google Scholar] [CrossRef]
  48. Li, Y.; Lin, S.; Ren, X.; Mi, H.; Zhang, P.; Sun, L.; Deng, L.; Gao, Y. One-step rapid in-situ synthesis of nitrogen and sulfur co-doped three-dimensional honeycomb-ordered carbon supported PdNi nanoparticles as efficient electrocatalyst for oxygen reduction reaction in alkaline solution. Electrochim. Acta 2017, 253, 445–454. [Google Scholar] [CrossRef]
  49. Sun, L.; Liao, B.; Ren, X.; Li, Y.; Zhang, P.; Deng, L.; Gao, Y. Ternary PdNi-based nanocrystals supported on nitrogen-doped reduced graphene oxide as highly active electrocatalysts for the oxygen reduction reaction. Electrochim. Acta 2017, 235, 543–552. [Google Scholar] [CrossRef]
  50. Niu, L.; Meng, K.; Song, J.; Liu, Y.; Jia, Z.; Wang, Y.; Tian, L.; Qi, T. Pd–Ni ellipsoidal nano-alloys with excellent catalytic performance for oxygen reduction reaction. Int. J. Hydrogen Energy 2024, 51, 234–244. [Google Scholar] [CrossRef]
  51. Celik, C.; San, F.G.B.; Sarac, H.I. Effects of operation conditions on direct borohydride fuel cell performance. J. Power Sources 2008, 185, 197–201. [Google Scholar] [CrossRef]
  52. Atwan, M.H.; Macdonald, C.L.B.; Northwood, D.O.; Gyenge, E.L. Colloidal Au and Au-alloy catalysts for direct borohydride fuel cells: Electrocatalysis and fuel cell performance. J. Power Sources 2006, 158, 36–44. [Google Scholar] [CrossRef]
  53. Hjelm, R.M.E.; Lafforgue, C.; Atkinson, R.W.; Garsany, Y.; Swider-Lyons, K. Impact of the anode catalyst layer design on the performance of H2O2-direct borohydride fuel cells. J. Electrochem. Soc. 2019, 166, F1218–F1228. [Google Scholar] [CrossRef]
  54. Wang, Z.; Mandal, M.; Sankarasubramanian, S.; Huang, G.; Kohl, P.A.; Ramani, V.K. Influence of water transport across microscale bipolar interfaces on the performance of direct borohydride fuel cells. ACS Appl. Energy Mater. 2020, 3, 4449–4456. [Google Scholar] [CrossRef]
  55. Ko, Y.; Lombardo, L.; Li, M.; Pham, T.H.M.; Yang, H.; Züttel, A. Selective borohydride oxidation reaction on nickel catalyst with anion and cation exchange ionomer for high-performance direct borohydride fuel cells. Adv. Energy Mater. 2022, 12, 2103539. [Google Scholar] [CrossRef]
  56. Wang, Z.; Sankarasubramanian, S.; Ramani, V. Reactant-transport engineering approach to high-power direct borohydride fuel cells. Cell Rep. Phys. Sci. 2020, 1, 100084. [Google Scholar] [CrossRef]
  57. Yin, X.; Zhu, K.; Ye, K.; Yan, J.; Cao, D.; Zhang, D.; Yao, J.; Wang, G. High efficiency N/C foam supported Pd electrode for direct sodium borohydride-hydrogen peroxide fuel cell. J. Power Sources 2022, 541, 231704. [Google Scholar] [CrossRef]
Figure 1. Preparation steps for PdNi@NG nanocomposites.
Figure 1. Preparation steps for PdNi@NG nanocomposites.
Sustainability 16 07469 g001
Figure 2. Schematic diagram of the assembled direct borohydride fuel cell (DBFC).
Figure 2. Schematic diagram of the assembled direct borohydride fuel cell (DBFC).
Sustainability 16 07469 g002
Figure 3. (a) TEM image of PdNi_15@NG and (b) N2-adsorption/desorption isotherms (i.e., volume of adsorbed N2 gas, Va, at standard temperature and pressure (STP) versus relative gas pressure, p/p0) of Ni@NG, PdNi_5@NG, PdNi_10@NG, and PdNi_15@NG electrocatalysts.
Figure 3. (a) TEM image of PdNi_15@NG and (b) N2-adsorption/desorption isotherms (i.e., volume of adsorbed N2 gas, Va, at standard temperature and pressure (STP) versus relative gas pressure, p/p0) of Ni@NG, PdNi_5@NG, PdNi_10@NG, and PdNi_15@NG electrocatalysts.
Sustainability 16 07469 g003
Figure 4. (a) The XPS elemental survey spectra of PdNi_15@NG, with (b) the high-resolution C 1s XPS spectrum, (c) the high-resolution O 1s XPS spectrum, (d) the high-resolution N 1s XPS spectrum, (e) the high-resolution Ni 2p XPS spectrum, and (f) the high-resolution Pd 3d XPS spectrum.
Figure 4. (a) The XPS elemental survey spectra of PdNi_15@NG, with (b) the high-resolution C 1s XPS spectrum, (c) the high-resolution O 1s XPS spectrum, (d) the high-resolution N 1s XPS spectrum, (e) the high-resolution Ni 2p XPS spectrum, and (f) the high-resolution Pd 3d XPS spectrum.
Sustainability 16 07469 g004
Figure 5. (a) The Raman spectra, (b) the FTIR spectra, and (c) the XRD patterns of Ni@NG, PdNi_5@NG, PdNi_10@NG, and PdNi_15@NG electrocatalysts.
Figure 5. (a) The Raman spectra, (b) the FTIR spectra, and (c) the XRD patterns of Ni@NG, PdNi_5@NG, PdNi_10@NG, and PdNi_15@NG electrocatalysts.
Sustainability 16 07469 g005
Figure 6. CVs of PdNi_5@NG (a), PdNi_10@NG (b), and PdNi_15@NG (c) in 2 M NaOH and 0.03 M NaBH4 + 2 M NaOH at 50 mV s−1 and 25 °C. (d) Direct comparison of forward scans of PdNi_5@NG, PdNi_10@NG, and PdNi_15@NG.
Figure 6. CVs of PdNi_5@NG (a), PdNi_10@NG (b), and PdNi_15@NG (c) in 2 M NaOH and 0.03 M NaBH4 + 2 M NaOH at 50 mV s−1 and 25 °C. (d) Direct comparison of forward scans of PdNi_5@NG, PdNi_10@NG, and PdNi_15@NG.
Sustainability 16 07469 g006
Figure 7. Anodic scans (ac) of the PdNi@NG electrocatalysts in 0.03 M NaBH4 + 2 M NaOH at 25 °C at different scan rates, ν, and (d) the corresponding jp vs. ν1/2 plots.
Figure 7. Anodic scans (ac) of the PdNi@NG electrocatalysts in 0.03 M NaBH4 + 2 M NaOH at 25 °C at different scan rates, ν, and (d) the corresponding jp vs. ν1/2 plots.
Sustainability 16 07469 g007
Figure 8. CVs at 10 mV s−1 of (a) Ni@NG and (bd) PdNi@NG electrocatalysts in 0.1 M KOH in the presence and absence of O2.
Figure 8. CVs at 10 mV s−1 of (a) Ni@NG and (bd) PdNi@NG electrocatalysts in 0.1 M KOH in the presence and absence of O2.
Sustainability 16 07469 g008
Figure 9. LSVs with electrode rotation of Ni@NG (a), PdNi_5@NG (b), PdNi_10@NG (c), and PdNi_15@NG (d) in 0.1 M KOH at 10 mV s−1 and corresponding K–L plots (a′,b′,c′,d′) at different potentials. The inset in (d′) shows the determined number of transferred electrons at different potentials for ORR at PdNi_15@NG.
Figure 9. LSVs with electrode rotation of Ni@NG (a), PdNi_5@NG (b), PdNi_10@NG (c), and PdNi_15@NG (d) in 0.1 M KOH at 10 mV s−1 and corresponding K–L plots (a′,b′,c′,d′) at different potentials. The inset in (d′) shows the determined number of transferred electrons at different potentials for ORR at PdNi_15@NG.
Sustainability 16 07469 g009
Figure 10. Comparison of (a) the ORR RDE voltammograms at 10 mVs−1 and 1600 rpm using prepared electrocatalysts and the corresponding (b) K–L plots at −0.5 V and (c) Tafel plots.
Figure 10. Comparison of (a) the ORR RDE voltammograms at 10 mVs−1 and 1600 rpm using prepared electrocatalysts and the corresponding (b) K–L plots at −0.5 V and (c) Tafel plots.
Sustainability 16 07469 g010
Figure 11. Cell polarization and corresponding power density curves for a DBFC using PdNi_15@NG as anode and cathode electrocatalyst in 1 M NaBH4 + 2 M NaOH anolyte and O2-saturated 1 M NaOH catholyte at temperatures ranging from 25 to 55 °C.
Figure 11. Cell polarization and corresponding power density curves for a DBFC using PdNi_15@NG as anode and cathode electrocatalyst in 1 M NaBH4 + 2 M NaOH anolyte and O2-saturated 1 M NaOH catholyte at temperatures ranging from 25 to 55 °C.
Sustainability 16 07469 g011
Table 1. The metal loading in the electrocatalysts as determined by ICP-OES analysis.
Table 1. The metal loading in the electrocatalysts as determined by ICP-OES analysis.
CatalystMetal Loading in the Electrocatalysts/wt.%
PdNi
Ni@NG-92.28 ± 0.14
PdNi_5@NG0.09 ± 0.00556.94 ± 0.23
PdNi_10@NG0.12 ± 0.0153.32 ± 0.21
PdNi_15@NG0.27 ± 0.0558.06 ± 0.26
Table 2. Pore structure parameters of samples.
Table 2. Pore structure parameters of samples.
CatalystSBET/m2 g−1Total Pore Volume/cm3 g−1Average Pore Diameter/nm
Ni@NG970.062.6
PdNi_5@NG710.084.3
PdNi_10@NG400.099.0
PdNi_15@NG60.321.5
Table 3. Comparison of the ORR performance of different electrocatalysts in 0.1 M KOH.
Table 3. Comparison of the ORR performance of different electrocatalysts in 0.1 M KOH.
CatalystEonset/V
vs. RHE
E1/2/V
vs. RHE
jd/
mA cm−2
nTafel Slope/V dec−1Source
Ni@NG0.690.60−1.162.3–2.50.100This work
PdNi_5@NG0.710.63−0.711.9–20.135This work
PdNi_10@NG0.730.61−1.412.10.108This work
PdNi_15@NG0.800.67−1.533.2–3.60.157This work
PdNi-4500.950.87−0.424.2–4.3[47]
PdNi-NS/C0.99 0.93 −0.213.70.080[48]
PdNiCu/NG0.87 0.761.794[49]
PdNi EN/CB0.990.920.194[50]
Table 4. Operational parameters of the DBFC employing the PdNi_15@NG electrocatalysts as the anode and cathode.
Table 4. Operational parameters of the DBFC employing the PdNi_15@NG electrocatalysts as the anode and cathode.
Temperature/°CE at Peak Power Density/Vj at Peak Power Density/mA cm−2Peak Power Density/
mW cm−2
Specific Peak Power Density
/kW gPd−1
250.50147.13.46
350.45187.93.85
450.45208.84.29
550.50199.54.63
Table 5. The performance of DBFCs using different Pd-based electrocatalysts.
Table 5. The performance of DBFCs using different Pd-based electrocatalysts.
Anode (mgPd cm−2)Cathode (mg cm−2)AnolyteCatholyteP/kW gPd−1T/°CSource
PdNi_15@NG (0.002)PdNi_15@NG1 M NaBH4 +
2 M NaOH
O2 + 1 M NaOH3.4625This work
Pd50Cu50/C
(0.313)
Pt/C
(0.5)
1 M NaBH4 +
6 M NaOH
O2 flowrate:
200 mL min−1
0.31360[12]
Pd/C
(1.08)
Pt/C
(0.3)
1 M NaBH4 +
20 wt.%NaOH
Air flowrate:
30 mL min−1
0.01825[51]
Au-Pd/C(1:1)
(1.75)
Pt/C
(4)
2 M NaBH4 +
2 M NaOH
O2 flowrate:
200 mL min−1
0.027 60[52]
53 wt.% Pd/C
(0.31)
50 wt.% Pt/C
(0.31)
0.1 M NaBH4 +
1 M NaOH
0.4 M H2O2 +
1 M H2SO4
0.88725[53]
Pd/C
(3)
Pt/C
(3)
1.5 M NaBH4 +
3 M KOH
15 wt.% H2O2 +
1.5 M H2SO4
0.20325[54]
NiF@AEI + Pd/C (0.5)Pt/C
(1)
1.5 M NaBH4 +
3 M NaOH
15 wt.% H2O2 +
1.5 M H2SO4
0.89880[55]
Pd/C (3) + NiPt/C
(3)
1.5 M NaBH4 +
3 M KOH
15 wt.% H2O2 +
1.5 M H2SO4
0.29770[56]
Pd@N/C-8 foam (0.15)Pd@N/C-8 foam
(0.15)
0.3 M NaBH4 +
2 M NaOH
1.6 M H2O2 +
2 M H2SO4
0.91360[57]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Balčiūnaitė, A.; Elessawy, N.A.; Šljukić, B.; Toghan, A.; Al-Hussain, S.A.; Gouda, M.H.; Youssef, M.E.; Santos, D.M.F. Effective Fuel Cell Electrocatalyst with Ultralow Pd Loading on Ni-N-Doped Graphene from Upcycled Water Bottle Waste. Sustainability 2024, 16, 7469. https://doi.org/10.3390/su16177469

AMA Style

Balčiūnaitė A, Elessawy NA, Šljukić B, Toghan A, Al-Hussain SA, Gouda MH, Youssef ME, Santos DMF. Effective Fuel Cell Electrocatalyst with Ultralow Pd Loading on Ni-N-Doped Graphene from Upcycled Water Bottle Waste. Sustainability. 2024; 16(17):7469. https://doi.org/10.3390/su16177469

Chicago/Turabian Style

Balčiūnaitė, Aldona, Noha A. Elessawy, Biljana Šljukić, Arafat Toghan, Sami A. Al-Hussain, Marwa H. Gouda, M. Elsayed Youssef, and Diogo M. F. Santos. 2024. "Effective Fuel Cell Electrocatalyst with Ultralow Pd Loading on Ni-N-Doped Graphene from Upcycled Water Bottle Waste" Sustainability 16, no. 17: 7469. https://doi.org/10.3390/su16177469

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