Sumera Akrama,
Shabbir Hussain*a,
Muhammad Arif
b,
Mirza Haider Alic,
Muhammad Tariq
d,
Abdur Raufe,
Khurram Shahzad Munawar
fg,
Hamad M. Alkahtani
h,
Amer Alhaj Zeni and
Syed Adnan Ali Shahjk
aInstitute of Chemistry, Khwaja Fareed University of Engineering and Information Technology, Rahim Yar Khan 64200, Pakistan. E-mail: shabbir.hussain@kfueit.edu.pk; shabchem786@gmail.com
bInstitute of Chemical and Environmental Engineering, Khwaja Fareed University of Engineering and Information Technology, Rahim Yar Khan 64200, Pakistan
cDepartment of Chemistry, Lahore Garrison University, Lahore, Pakistan
dInstitute of Chemical Sciences, Bahauddin Zakariya University, Multan, Pakistan
eDepartment of Chemistry, University of Sahiwal, Sahiwal, Pakistan
fInstitute of Chemistry, University of Sargodha, 40100, Pakistan
gDepartment of Chemistry, University of Mianwali, 42200, Pakistan
hDepartment of Pharmaceutical Chemistry, College of Pharmacy, King Saud University, P. O. Box 2457, Riyadh 11451, Saudi Arabia
iChemistry & Forensics Department, Nottingham Trent University, Clifton Campus, Nottingham Ng11 8NS, UK
jFaculty of Pharmacy, Universiti Teknologi MARA Cawangan Selangor Kampus Puncak Alam, Bandar Puncak Alam, Selangor D. E. 42300, Malaysia
kAtta-ur-Rahman Institute for Natural Product Discovery (AuRIns), Universiti Teknologi MARA Cawangan Selangor Kampus Puncak Alam, Bandar Puncak Alam, Selangor D. E. 42300, Malaysia
First published on 1st May 2025
In this research, we synthesized (Co3O4)aq and (Co3O4)et nanoparticles (NPs) utilizing aqueous and ethanolic extracts, respectively, of Psidium guajava leaves. The biosynthesized NPs were sonicated with reduced graphene oxide (rGO) to produce rGO@(CO3O4)aq and rGO@(Co3O4)et nanocomposites (NCs) and their respective calcined (700 °C) products rGO@(CO3O4)aqc and rGO@(Co3O4)etc. The nanomaterials (NMs) were characterized through XRD, FTIR, UV-visible spectroscopy, SEM, TGA, and DSC analyses. They exhibited crystallite sizes of 10–15.4 nm and band gaps of 5.1–5.9 mV. Their surfaces were coated with organic moieties from plant extracts. TGA and DSC analyses showed the endothermic loss of moisture and exothermic evolution of organic contents. SEM images revealed the rough and porous surfaces of NPs, making them efficient catalysts for water splitting. Linear swap voltammetry (LSV) measurements for the oxygen evolution reaction (OER) and hydrogen evolution reaction (HER), Tafel slopes and double layer capacitance (Cdl) values reflected a decrease in electrocatalytic water splitting efficiency in the following order: rGO@(Co3O4)aq > (Co3O4)aq > rGO@(Co3O4)aqc and rGO@(Co3O4)et > (Co3O4)et > rGO@(Co3O4)etc. Each aqueous extract-derived nanomaterial was electrocatalytically more active than its respective ethanolic extract-derived counterpart. Moreover, the non-calcined rGO decorated Co3O4 products showed superior electrocatalytic performance compared with their calcined counterparts and therefore, can be recommended as the best choices for electrocatalytic water splitting applications.
Earlier studies have reported that transition metal oxides display promising properties such as high specific surface areas and diverse morphologies. However, their practical applications are limited by certain drawbacks that can be addressed by integrating carbon-based materials, especially graphene oxide (GO) with metal oxides (MOs). Such synergistically integrated nanocomposites combine the good electrochemical potential and rich redox activity of metal oxides with the ultra-thin structure, large specific surface area and exceptional thermal/electrical conductivity of GO. GO–MO hybrid materials exhibit superior specific capacitance and cyclic stability in ultracapacitors compared with their individual components and offer excellent potential for advancing next-generation supercapacitor materials.6 They exhibit large surface areas, high chemical stability, excellent thermal and electrical conductivity, and have attracted significant attention in energy conversion processes like oxygen reduction reactions (ORR) and hydrogen evolution reactions (HER) and in energy storage devices like supercapacitors, batteries,7 solar cells and fuel cells.8 Among the carbon-based materials, graphene and carbon nanotubes (CNT) are regarded as the new-generation and state-of-the-art nano-reinforcement for metals owing to their outstanding multifunctional features, extraordinary mechanical properties, and unique nanostructures.9 However, graphene has attracted more attention than CNTs owing to its higher aspect ratios, 2D flat geometry, unique surface texture (capability to interlock mechanically with the matrix), cost-effective production,10 better surface properties (under ambient conditions) and higher selectivity against interferences.11 Currently, GO has emerged as a versatile substance with its outstanding properties in energy conversion and storage technologies.8 GO–MO hybrids display improved surface area compared with their individual constituents, resulting in better charge separation properties, high and selective adsorption capacity towards metal ions and organic species, and photocatalytic degradation of pollutant dyes and pathogens. They are also effective in energy storage, water purification, antibacterial applications, controlled drug release and selective destruction of cancerous cells.12 They exhibit improved conductivity, stability, and reactivity by leveraging the strengths of GO and metal oxides, making them promising candidates for advanced materials in numerous technological domains, including electronic devices, catalysts, sensors and energy storage devices. GO contributes to the improved surface functionality, mechanical strength and excellent conductivity, whereas metal oxides are attributed to specific chemical and electronic properties.8 GO finds significance due to its increased polarity and compatibility with other nanomaterials, making it an important component in composite electrodes.6 Reduced graphene oxide (rGO) provides a flat and broad surface with several functional groups and defects, which increases its dispersion and integration with metal oxides. The increased roughness factor and surface area significantly promotes its electrochemical potential and electrocatalytic efficiency.13 rGO-metal oxide nanocomposites combine the pseudocapacitive properties of metal oxides with the conductivity of rGO14 and are highly suitable candidates for supercapacitors due to their excellent mechanical behavior, good chemical stability, superior electrical conductivity and high surface area.15 They find multifaceted applications in designing and fabricating smart materials due to their unique photochemical-, photocatalytic-, sensing-, mechanical-, electrical-, optical- and energy-storing capacities.16
Cobalt oxide-rGO nanocomposites were earlier reported as electrocatalysts for OER,17 electrode materials for supercapacitors,18 anode materials for lithium ion batteries,19 advanced multifunctional microwave absorbers,20 acetone sensors21 and sensitive electrochemical detectors of trace Pb(II) ions in environmental samples.22 Cobalt NPs have a high surface area to volume ratio and display excellent catalytic activity,23 while rGO exhibits good carrier transportation, excellent thermal and chemical stability, electrical conductivity, hydrophobicity, and safety, high mechanical strength, substantial specific surface area,24 and good durability and performance, for advanced materials.25 Under moderate hydrothermal conditions, the surface contact between graphene oxide (GO) sheets and Co2+ ions may be altered by deoxygenating a few layers of GO.26 The presence of oxygen-containing functional groups on rGO surface facilitates its better bonding with Co3O4 NPs. These groups act as anchoring sites, helping in the uniform distribution of Co3O4 NPs and improving the composite's stability and performance. On the other hand, CNTs have fewer surface functional groups, making it harder for them to interact with metal oxides in a similar way. The choice of rGO over GO for synthesizing composites stems from its superior electrical conductivity, higher mechanical stability, and improved chemical resistance, which are crucial for high-performance applications.27 Due to the greater portion of sp3 hybridized carbons linked with the oxygen-containing moieties, GO is typically insulating and displays a very high sheet resistance. The sheet resistance of GO is greatly lowered after its reduction into rGO, hence converting it into a semiconductor or even into a graphene-like semimetal.12
There are numerous physical and chemical processes for creating NPs, but green synthetic pathways are environmentally friendly and the most suitable methods28 because they reduce the use of harmful chemicals, solvents, and other synthetic agents and minimize the toxic environmental effects associated with chemical synthesis.29 Green synthesis has received substantial attention in recent years, especially in the field of materials science30 and researchers are constantly exploring new and innovative nano-synthetic methods via the use of renewable energy sources such as plant extracts.31 Plants are rich sources of antioxidants including phenolics, flavonoids, fatty acids, etc.;32 therefore, they can be employed as reducing agents for the conversation of metal salts into their respective NPs.
In the present study, we have employed the aqueous and ethanolic extracts of Psidium guajava leaves to synthesize cobalt oxide (Co3O4) NPs and then their nanocomposites with rGO. The investigated biosynthesis can be attributed to the diverse phytochemical profile of P. guajava which includes flavonoids, polyphenols, terpenoids, and tannins.33 In particular, flavonoids and polyphenols are known for their strong reducing and capping abilities, allowing for the controlled growth and stabilization34 of cobalt ions during the production of NPs. The biosynthesized nanomaterials were analyzed by XRD studies, FTIR analysis, UV-visible spectroscopy, SEM studies, TGA and DSC analyses. They were also tested for their electrocatalytic water splitting potential by linear sweep voltammetry (LSV), HER and OER measurements and their surface areas were compared in terms of double-layer capacitance (Cdl). This research aimed to comprehensively explore the diverse and ever-expanding utilization of cobalt oxide nanoparticles for water splitting, highlighting their immense potential for driving innovation and addressing contemporary challenges. The main objective was to develop an electrode with enhanced electrochemical performance through combination of cobalt oxide with graphene material.
Powder XRD analysis was performed using a PANalytical X'Pert Pro X-ray diffractometer. Other analytical instruments include Shimadzu FTIR-8400 Fourier transform infrared spectrometer, CE 7200 double-beam UV-visible spectrophotometer for UV-visible spectroscopy and an Emcrafts Cube 1100 for SEM analysis. Thermogravimetric analysis (TGA) and differential scanning calorimetry (DSC) were performed from ambient temperature to 1000 °C using a TA Instruments Discovery 650 SDT simultaneous thermal analyzer, with a heating rate of 10 °C per minute under a 99.999% nitrogen atmosphere and an average flow rate of 50 mL min−1 was maintained. An electrochemical workstation (potentiostat, CHI 760 E, CH Instrument Co., USA) was used to conduct the electrochemical studies.
The electrochemical experiments were conducted using a standard three-electrode system, with nickel foam as the working electrode, a platinum wire as the counter electrode and an Ag/AgCl electrode filled with saturated KCl as a reference electrode which was calibrated to the reversible hydrogen electrode (RHE) in a 1 M KOH-saturated electrolyte. To activate the catalyst-coated electrodes, 20 cycles were run by performing cyclic voltammetry (CV) at a 0.1 V s−1 scan rate within the 0.2–0.8 V vs. Ag/AgCl potential range. Additional CVs were recorded at various scan rates (100, 80, 60 and 40 mV s−1) to determine the double-layer capacitance (Cdl).
Five g of P. guajava leaf powder was mixed with 250 mL of ethanol, stirred at 40 °C for 30 min, followed by filtration to obtain the ethanolic extract (filtrate), which was employed for the synthesis of (Co3O4)et NPs.
The same procedure was followed for the green synthesis of (Co3O4)et NPs except that the ethanolic extract of P. guajava leaves was used instead of the aqueous extract. Fig. 2 summarizes the whole biosynthetic route of (Co3O4)aq and (Co3O4)et NPs.
0.01 g of rGO suspension in 50 mL of distilled water was sonicated for 150 min. Then, 0.1 g of (Co3O4)aq NPs were added and the mixture was subjected to another 3.5 h sonication until a homogeneous mixture was obtained. The as formed nanocomposite was separated by centrifugation (10000 rpm) for 20 minutes and washed thrice with distilled water, followed by centrifugation each time and dried at 60 °C in an incubator/oven to produce solid rGO@(Co3O4)aq. Finally, rGO@(Co3O4)aq was calcined at 700 °C for 120 min in a muffle furnace to produce rGO@(Co3O4)aqc.
The same methodology (discussed above) was followed for the biosynthesis of rGO@(Co3O4)et nanocomposites from (Co3O4)et and rGO. The resulting rGO@(Co3O4)et product was calcined at 700 °C for 120 min to produce rGO@(Co3O4)etc. Fig. 3 displays the synthetic route for the synthesis of rGO decorated Co3O4 nanocomposites.
To make the binder, 0.2 g of PVDF was added to 5 mL of N-methyl-2-pyrrolidone (NMP), and the mixture was agitated for 24 h.
To prepare the electrode, a China dish containing 0.08 g of a sample's material was utilized. Then one drop of the binder was combined with 0.01 grams of activated carbon and gently blended. The resulting mixture was subsequently applied twice to both sides of the activated nickel foam strips using a small painting brush. Finally, the coated strips were oven-dried for 12 h at 40 °C.37
The 2θ values of 10.53°, 18.42°, 31.35°, 36.64°, 38.01°, 43.87°, 54.87°, 58.01°, 64.31° and 67.62° in rGO@(Co3O4)aq and of 10.66°, 17.29°, 19.50°, 28.53°, 34.01°, 35.37°, 37.78°, 43.98°, 58.01°, 54.23°, 58.45° and 66.51° in rGO@(Co3O4)et agree with the incorporation of rGO into Co3O4 in these two nanostructures. The rGO peak is displayed at 10.53° and 10.66° in rGO@(Co3O4)aq and rGO@(Co3O4)et, respectively. A peak at 43.87° and 43.98° in rGO@(Co3O4)aq and rGO@(Co3O4)et, respectively can been assigned to metallic cobalt. The diffraction peaks appeared at 2θ values of 11.20°, 18.43°, 30.38°, 36.32°, 43.21°, 48.41°, 53.91°, 58.91° and 64.31° in rGO@(Co3O4)aqc and of 11.26°, 18.29°, 30.74°, 36.97°, 44.60°, 48.22°, 54.44°, 58.86° and 64.49° in rGO@(Co3O4)etc. Both of these calcined samples i.e., rGO@(Co3O4)aqc and rGO@(Co3O4)etc represent the purer Co3O4 phase compared with their non-calcined counterparts i.e., rGO@(Co3O4)aq and rGO@(Co3O4)et, respectively. A peak at 11.20° in rGO@(Co3O4)aqc and at 11.66° in rGO@(Co3O4)etc represents the incorporation of reduced graphene oxide in these two nanomaterials.
The Debye–Scherrer formula, D = Kλ/βcos(θ), was used to calculate average crystallite sizes of the nanomaterials. In this equation, K represents the shape factor, λ is the wavelength of X-ray radiations, β denotes the full width at half maximum (FWHM) of the diffraction peak, θ is the Bragg angle, and D indicates the calculated crystallite sizes. The aqueous extract-derived (Co3O4)aq possessed the smaller crystallite size (10 nm) compared with its ethanolic extract-derived counterpart (15.4 nm). Actually, the solvent used for extracting bioactive compounds from P. guajava leaves (like water vs. ethanol) can affect the types and quantities of bioactive compounds present. The results of our study demonstrate that the aqueous extract of P. guajava leaves influences the nucleation and growth dynamics of NPs in different ways compared with the ethanolic extract to produce different sized NPs in both cases.39 rGO@(Co3O4)aq and rGO@(Co3O4)et have shown an average crystallite size of 12 and 11.57 nm, respectively which was further lowered to 11.5 nm in both the calcined nanocomposites i.e., rGO@(Co3O4)aqc and rGO@(Co3O4)etc. The results of our study clearly demonstrate that the average crystallite sizes of biosynthesized nanomaterials vary depending upon the nature of plant extraction solvent, the doping of Co3O4 with rGO and the calcination temperature.
![]() | ||
Fig. 5 FT-IR spectra: (a) (Co3O4)aq, rGO@(Co3O4)aq and rGO@(Co3O4)aqc; (b) (Co3O4)et, rGO@(Co3O4)et and rGO@(Co3O4)etc. |
![]() | ||
Fig. 6 UV-visible spectra: (a) (Co3O4)aq, rGO@(Co3O4)aq and rGO@(Co3O4)aqc; (b) (Co3O4)et, rGO@(Co3O4)et and rGO@(Co3O4)etc. |
Band gap values, indicative of the energy required for electronic transitions within the materials, decreased from 5.4 eV in (Co3O4)aq to 5.3 and 5.1 eV in its nanocomposites rGO@(Co3O4)aq and rGO@(Co3O4)aqc, respectively. However, there was a significant rise in the band gap of (Co3O4)et (5.3 eV) to 5.9 and 5.8 eV in its nanocomposites rGO@(Co3O4)et and rGO@(Co3O4)etc, respectively. The variations in band gaps may be owing to the different nature of plant coatings (aqueous and ethanolic extracts) on the surfaces of synthesized nanomaterials. The smaller and higher bandgap materials have their own advantages and disadvantages. These values provide insights into the material's optical properties and electronic structures, which are crucial for various applications. In water splitting applications, materials with smaller band gap energies are generally more effective.44
![]() | ||
Fig. 7 SEM images of (a) (Co3O4)aq, (b) (Co3O4)et, (c) rGO@(Co3O4)aq, (d) rGO@(Co3O4)et, (e) rGO@(Co3O4)aqc, and (f) rGO@(Co3O4)etc. |
![]() | ||
Fig. 8 TGA and DTA curve of (Co3O4)aq, rGO@(Co3O4)aq, rGO@(Co3O4)aqc, (Co3O4)et, rGO@(Co3O4)et, and rGO@(Co3O4)etc. |
The obtained results (Fig. 8) reflect the thermal instabilities of the synthesized nanomaterials. According to literature, the initial weight loss at 25–150 °C in the TGA curve, concomitant with an endothermic peak in the DSC curve, is attributed to the evaporation of residual moisture from the investigated samples.48 The DSC curves of (Co3O4)aq, rGO@(Co3O4)aq and rGO@(Co3O4)aqc show endothermic peaks at 92, 82 and 170 °C, respectively, corresponding to the evaporation of adsorbed water from the surfaces of these nanostructures.49 TGA curves also depict the gradual loss of mass corresponding to the exothermic peaks in DSC curves at 380, 378, 130, 163, 136 and 117 °C in (Co3O4)aq, rGO@(Co3O4)aq, rGO@(Co3O4)aqc, (Co3O4)et, rGO@(Co3O4)et, and rGO@(Co3O4)etc, respectively; these peaks demonstrate the exothermic evolution of residual organic moieties that are present as plant coatings on the surfaces of synthesized nanoparticles. An additional exothermic peak at 360 °C was also displayed in the DSC curve of rGO@(Co3O4)aqc. Conclusively, TGA/DSC analysis shows endothermic loss of moisture and organic contents from the NMs.
![]() | ||
Fig. 9 LSV OER curves (left) and corresponding Tafel slopes (right) of (Co3O4)aq, rGO@(Co3O4)aq, rGO@(Co3O4)aqc, (Co3O4)et, rGO@(Co3O4)et, and rGO@(Co3O4)etc. |
Overall, the OER electrocatalytic potential of the biosynthesized nanomaterials was reduced in the following order:
rGO@(Co3O4)aq > (Co3O4)aq > rGO@(Co3O4)aqc > rGO@(Co3O4)et > (Co3O4)et > rGO@(Co3O4)etc |
The obtained results clarify that all the aqueous extract-based nanomaterials show better OER catalytic performance compared with their ethanolic extract based counterparts; this is mainly because of the differences in the nature of plant coatings (aqueous or ethanolic extract based) on the surfaces of investigated nanomaterials. Moreover, rGO decorated non-calcined rGO@(Co3O4)aq and rGO@(Co3O4)et nanocomposites showed superior electrocatalytic water splitting performance compared with their calcined rGO@(Co3O4)aqc and rGO@(Co3O4)etc counterparts, respectively. This is due to more rough and porous morphologies of the non-calcined materials compared with their calcined counterparts as discussed in Section 3.5. Calcination decreases the roughness and porosity with a consequent lowering of the electrocatalytic water splitting performance. The obtained results also clarify that the HER electrocatalytic performance of (Co3O4)aq and (Co3O4)et is significantly improved after forming their nanocomposites with GO i.e., rGO@(Co3O4)aq and rGO@(Co3O4)et, respectively but this performance is significantly lowered after calcination at 700 °C in rGO@(Co3O4)aqc and rGO@(Co3O4)etc.
The Tafel slope is a crucial parameter in electrochemistry, particularly when evaluating the performance of catalysts for OER. It essentially describes the relationship between the overpotential and the current density in an electrochemical reaction, such as OER. The Tafel slope indicates how sensitive the current density is to changes in overpotential. Generally, a smaller Tafel slope indicates a more efficient catalyst, as it means less energy is required to drive a significant increase in current density. The electrocatalytic kinetics of the biosynthesized samples were compared on the basis of their Tafel slopes. The rGO@(Co3O4)aq nanocomposite displayed the lowest Tafel slope (31 mV dec−1) compared with the remaining aqueous extract-based materials i.e., (Co3O4)aq (36 mV dec−1) and rGO@(Co3O4)aqc (43 mV dec−1), verifying that rGO@(Co3O4)aq demonstrates the most efficient OER kinetics. Among the ethanolic extract-based nanocomposites, the highest OER kinetics was displayed by rGO@(Co3O4)et (Tafel slope = 47 mV dec−1) compared with (Co3O4)et (51 mV dec−1) and rGO@(Co3O4)etc (55 mV dec−1) (Fig. 9).
The potential-dependent Tafel slopes for metal oxides can also be predicted from microkinetic models. A theoretical study on RuO2 demonstrated that the rate-determining step is O–O bond formation (*O + H2O → *OOH + H+ + e−) which determined the Tafel slope value of ∼39 mV dec−1 at a potential lower than ∼1.5 VRHE, where the active coordinatively unsaturated Ru sites (*) is filled with *OH.50 Scott et al.51 observed a Tafel slope of approximately 25 mV dec−1 at low overpotentials on Ru-based oxides. They proposed that this low Tafel slope could be attributed to the potential dependence of the coverage of surface species participating in the rate-determining step. This implies that the observed Tafel slope is influenced by the equilibrium coverage of intermediates at low potentials, leading to a distinct mechanistic pathway. Krasil'shchikov's path describes one of the most renowned mechanisms with the corresponding Tafel slopes (eqn (1)–(4)).
M + OH− ↔ MOH + e−, b = 120 mV dec−1 | (1) |
MOH + OH− ↔ MO− + H2O, b = 60 mV dec−1 | (2) |
MO− → MO + e−, b = 45 mV dec−1 | (3) |
2MO → 2M + O2, b = 19 mV dec−1 | (4) |
The rate determining step can be determined based on the Tafel slope. NiCo2O4 and IrO2 have Tafel slope values of 59 and 48 mV dec−1, respectively corresponding to their rate determining steps as displayed in steps 2 and 3, respectively. The smaller Tafel slope of IrO2 demonstrates reduced kinetic overpotential losses and can be owing to the rise of the bond strength for OH− adsorption on IrO2, which enhances the rate of the first electron reaction step (eqn (1)), consequently increasing the electrocatalytic kinetics. Moreover, the concentration of active sites and their contribution can be found from the changes in Tafel slopes. The reduced Tafel slope of IrO2 may also be owing to the rise of active sites and their active contribution.52 In our current study, the observed Tafel slope (31 mV dec−1) of rGO@(Co3O4)aq nanocomposite could indicate a mechanism involving specific adsorption processes, such as OH− adsorption, influencing the rate-determining step. This highlights the importance of considering surface coverage and adsorption phenomena when interpreting Tafel slopes in OER kinetics.53
![]() | ||
Fig. 10 LSV HER curves and corresponding Tafel slopes of (Co3O4)aq, rGO@(Co3O4)aq, rGO@(Co3O4)aqc, (Co3O4)et, rGO@(Co3O4)et, and rGO@(Co3O4)etc. |
The LSV HER experiments show that rGO@(Co3O4)aq, (Co3O4)aq and rGO@(Co3O4)aqc require a potential of 704, 708 and 714 mV, respectively to acquire a current density of 100 mA cm−2 and the corresponding Tafel slopes were 124, 134 and 170 mV dec−1, respectively. On the other hand, rGO@(Co3O4)et, (Co3O4)et and rGO@(Co3O4)etc achieved the same current density (100 mA cm−2) at the required potentials of 632, 635 and 700 mV, respectively with the corresponding Tafel slopes of 145, 153 and 161 mV dec−1, respectively.
A comparison of the Tafel slope values demonstrated that, rGO@(Co3O4)aq exhibited the lowest Tafel slope value of 124 mV dec−1 among the aqueous extract-derived materials, indicating that is has the most efficient HER kinetics compared with (Co3O4)aq (134 mV dec−1) and rGO@(Co3O4)aqc (170 mV dec−1). From the ethanolic extract-based nanomaterials, rGO@(Co3O4)et demonstrated the most favorable HER performance, requiring 632 mV to reach 100 mA cm−2, with the lowest Tafel slope of 145 mV dec−1.
Similar to the LSV OER results, the non-calcinated rGO@(Co3O4)aq and rGO@(Co3O4)et products show higher HER catalytic performance compared with their calcinated rGO@(Co3O4)aqc and rGO@(Co3O4)etc counterparts, respectively. rGO@(Co3O4)aq and rGO@(Co3O4)et show higher catalytic potential among their respective series of aqueous and ethanolic extract-derived products, respectively. However, a combined electro-catalytical comparison between all the products clarifies that aqueous extract-derived rGO@(Co3O4)aq and (Co3O4)aq products (Tafel slopes = 124 and 134 mV dec−1, respectively) are superior HER catalysts compared with their ethanolic extract-derived rGO@(Co3O4)et and (Co3O4)et counterparts, respectively (Tafel slopes = 145 and 153 mV dec−1, respectively). On the other hand, the ethanolic extract-derived calcined rGO@(Co3O4)etc product (Tafel slope = 161 mV dec−1) has a higher HER performance compared with its aqueous extract-derived calcined rGO@(Co3O4)aqc counterpart (Tafel slope = 170 mV dec−1). Overall, our study recommends that rGO@(Co3O4)aq and (Co3O4)aq are more useful HER electrocatalysts and can be applied more successfully for water splitting compared with the remaining nanomaterials.
NPs | Nanoparticles |
NCs | Nanocomposites |
NMs | Nanomaterials |
rGO | Reduced graphene oxide |
(Co3O4)aq and (Co3O4)et | Represent Co3O4 nanoparticles synthesized with aqueous and ethanolic extracts, respectively, of P. guajava leaves |
rGO@(CO3O4)aq and rGO@(Co3O4)et | Indicate the nanocomposites of (Co3O4)aq and (Co3O4)et, respectively, with reduced graphene oxide (rGO) |
rGO@(CO3O4)aqc and rGO@(Co3O4)etc | Represent the calcined products (at 700 °C) of rGO@(CO3O4)aq and rGO@(Co3O4)et, respectively |
Footnote |
† Electronic supplementary information (ESI) available: Reference XRD spectra of Co3O4 and cobalt hydroxide are shown in Fig. S1 and S2, respectively. Fig. S3 displays cyclic voltammograms of the (Co3O4)aq, rGO@(Co3O4)aq, rGO@(Co3O4)aqc, (Co3O4)et, rGO@(Co3O4)et and rGO@(Co3O4)etc samples. See DOI: https://doi.org/10.1039/d5ra00040h |
This journal is © The Royal Society of Chemistry 2025 |