Doungkamon Phihusut‡
a,
Joey D. Ocon‡ac,
Jae Kwang Leeab and
Jaeyoung Lee*ab
aElectrochemical Reaction and Technology Laboratory (ERTL), School of Environmental Science and Engineering, Gwangju Institute of Science and Technology (GIST), Gwangju 500-712, South Korea. E-mail: jaeyoung@gist.ac.kr
bERTL Center for Electrochemistry and Catalysis, RISE, GIST, Gwangju 500-712, South Korea
cLaboratory of Electrochemical Engineering (LEE), Department of Chemical Engineering, College of Engineering, University of the Philippines Diliman, Quezon City 1101, Philippines
First published on 16th July 2015
Low cost yet efficient water oxidation catalysts are crucial in making economically competitive water electrolyzers and secondary metal–air batteries. In this study, we demonstrate the optimized mild reduction of graphene oxide towards the synthesis of highly active and stable cobalt–graphene electrocatalysts for water oxidation. Contrary to the conventional use of fully reduced graphene oxide (RGO) as a composite material in electrocatalysis, our results suggest that the oxygen functional groups, which are retained during mild GO reduction, are crucial in the formation of cobalt oxalate (CoC2O4) microstructures. Gently reduced graphene oxide (gRGO) with a low degree of reduction results in CoC2O4/gRGO microrods with impressive water oxidation activity, reaching current densities 21.1% higher than conventional iridium oxide-based catalysts and 70.5% more than the unoptimized CoC2O4/gRGO catalysts. Mild reduction of GO favors the homogeneous formation of microstructures via the negatively-charged functional groups, which attract the positive Co ions and lead to a stronger chemical interaction between the two components. This work points towards investigating and reevaluating the role of the degree of GO reduction on graphene's contribution to the composition and catalytic activity of metal–graphene composites.
Aside from expensive noble metals, first-row transition metal spinels and perovskites have been investigated as OER electrocatalysts in alkaline condition.11–13 For instance, cobalt (Co)-based catalysts are well known to be good OER catalysts in alkaline media, in addition to having catalytic activity towards oxygen reduction reaction (ORR). Hence, a significant number of Co-based electrocatalysts have been suggested, varying mainly in its structure (i.e. thin films, nanowires, and nanoparticles) and chemical composition (i.e. metallic, oxides, hydroxides, and alloys).14–17 Of late, we first demonstrated the use of Co oxalate (CoC2O4) as a prospective precursor in making one-dimensional (1D) nanostructured electrodes with high catalytic activity for OER in alkaline media.18 Furthermore, another innovative concept is the application of hybrid materials based on Co and carbon (e.g. graphene, carbon nanofibers, and carbon nanotubes), in order to exploit the inherent advantages of these carbon materials in electrocatalysis (i.e. scalable production, inexpensive precursors, and good electronic conductivity).17,19–23 Combining both approaches, we recently developed a bifunctional catalyst, based on gently reduced graphene oxide incorporated into microstructured cobalt oxalate rods (CoC2O4/gRGO), showing excellent and stable electrocatalytic activity in both OER and ORR.24 While gRGO is inactive in OER, combining it with CoC2O4 significantly improves the OER stability and contributes to the ORR activity via the nitrogen functional groups, leading to a slightly higher ORR limiting current than platinum on carbon (Pt/C). Nevertheless, the relationship between the degree of graphene oxide reduction and its properties as substrate for the formation of CoC2O4 has yet to be understood.
In this paper, we report the facile synthesis of gRGOs at different reduction condition (i.e. time and temperature) and demonstrate its effectiveness for the incorporation with CoC2O4 as a highly active electrocatalyst for oxygen evolution from water. Using various physico-chemical characterization and electrochemical methods, we show that a strongly reduced GO is not required in the efficient catalysis of water oxidation using the cobalt–graphene composite.
X-Ray diffractograms of the gRGO samples were studied to look at the changes in the crystallinity. As shown in Fig. 2a, the diffraction patterns of gRGOs reduced at 90 °C showed a narrowing of the initially broad peak at around 23.9°, indicating an increase in the interlayer distance between the graphite oxide sheets.25 Expectedly, the small GO peak at around 12° disappeared as the reduction occurred. On the other hand, reduction proceeded at a lesser degree for gRGOs reduced at 25 °C, as proven also by XPS. Nonetheless, the XRD analysis proves that the GO was only mildly reduced, as exhibited by the absence of the sharp characteristic RGO peak usually found at 24°.
Fig. 2b and c display the representative CVs of the gRGOs treated at 90 °C and 25 °C. Evidently, the gRGOs reduced at 90 °C exhibited pseudo-capacitive characteristics in between −0.8 V and −0.2 V (vs. Hg/HgO), indicating the occurrence of an oxidation–reduction reaction pair.25 Pseudo-capacitive behaviour, however, was not observed in gRGOs reduced at 25 °C. Thus, the capacitive behaviour in the gRGO samples is a combination of the double-layer capacitance and pseudo-capacitance, in the case of gRGOs samples reduced at higher temperature.
The reduction of gRGO was also monitored using UV-Vis spectroscopy as seen in Fig. S1.† The absorption peaks of gRGOs are located at 265 nm, within the absorption region of fully reduced GO and graphene.27,28 The characteristic absorbance peak for the π–π* transition in pristine GO is located at 230 nm.29 During reduction of GO, however, the accompanying conjugation of the electronic structure shifted the absorbance peak to the higher wavelength region, near to that of graphene.27 The increasing absorbance at 265 nm with longer reduction times occurred with the solution color change from brown to black and with the increase in the amount of dispersed gRGO in the aqueous solutions.28,30 This is crucial in obtaining highly dispersible gRGO in water and in making CoC2O4/gRGO.
To further investigate the functional groups after reduction at low and high temperatures, FT-IR spectroscopy was performed. As shown in the FT-IR spectra of gRGOs in Fig. S2,† there are slight differences in the absorption bands at 1619 cm−1 and 3403 cm−1, denoting changes in the hydroxyl and carboxyl groups at different temperatures. The typical absorption bands of GO are located at 1060 cm−1 (C–O–C), 1395 cm−1 and 3403 cm−1 (O–H), 1619 cm−1 (CC) and 1739 cm−1 (C
O).27 Increasing the reduction temperature of the solution with diluted hydrazine, however, partially reduced the GO only.
On the other hand, the resistivities of gRGO paper samples, produced after drying the samples reduced at 90 °C at different reduction times, were measured using four-point probe analysis (Fig. 2d). The lowest resistivity was obtained for gRGO-90 °C-3 h (0.32 Ω cm), while high resistivities were measured for gRGOs reduced for a short time. The enhanced electrical conductivity of gRGOs signifies that the oxygen functional groups were removed and the defects were lost. In principle, removal of the oxygen functional groups during reduction leads to the reconnection of the original graphitic domains by forming new sp2 clusters to restore the long-range conjugation of GO.31 This could result to a significant rise in the electrical conductivity once a high number of the sp2 structures is achieved.32 The advantages of the improved conductivity have been widely used in electrocatalysis, as the electron transfer between the liquid–solid interfaces is crucial in accelerating the reactions. Indeed, this reasoning has been the consistent justification on why RGO is applied as composite material.33
In order to shed light on the electrochemical characteristics of gRGOs, the surface chemistry of gRGOs was investigated using X-ray photoelectron spectroscopy (XPS). The C1s spectra of gRGO were deconvoluted into five characteristic peaks consisting of the following bonds: sp2 (CC)/sp3 (C–C), hydroxyl and epoxide (C–O), carbonyl (C
O), carboxyl (O
C–O), and the shake-up satellite (π–π*).33,34 Using these results, we plotted the relative amounts of the different surface functional groups in the gRGO samples (Fig. 3a), except the C
C/C–C bonds, which constituted majority of the C1s signal (68.6%). C–O has the highest atomic percentage (>15 at%), while C
O and O
C–O were at 6.8 at% and 3.5 at%, respectively. Furthermore, a rising π–π* shake-up satellite peak was only observed for the reduction at 5 h, implying the presence of aromatics, conjugated systems or the electronic excitation of the π electron system on an aromatic ring by the photoelectrons.35–37
Additionally, we analyzed the functional groups in gRGOs reduced with the same set of reduction times (1 h, 3 h, and 5 h) but at 25 °C. Similar to the previous spectra, the signals were fitted with the same five characteristics peaks (Fig. 3c) and we plotted the relative amounts of the functional groups in Fig. 3d. The C–O (25.32 at%) groups exhibited the highest concentration, having several-folds higher content among the other functional groups. Similar to the results above, the presence of the π–π* shake-up satellite peak was observed for gRGO-25 °C-1 h, at a higher binding energy, reflecting the restoration of aromaticity after eliminating some oxygen atoms in the surface.38 Comparing the spectra of gRGOs reduced at 25 °C and 90 °C, the removal of C–O seemed to have influenced more the capacitance of gRGOs reduced at 90 °C. The difficulty in removing C–O groups at 25 °C is explained by the transformation and evolution mechanisms involved in the single-bonded C–O, consistent with previous studies.39 Various functional groups (i.e. C–OH, C–O–C, and –O) actually contributed to the C–O signal. Moreover, at low temperature reduction, the transformation of CO to C–O can also occur. The initial C–OH species are suggested to have been involved in the formation of phenols by sacrificing the C–O–C bonds.39 As mentioned above, pseudo-capacitance behaviour influenced in increasing the specific capacitance in gRGOs reduced at 90 °C. The increase in capacitance from zero to 1 h of reduction time at 90 °C could be attributed to the increase in the carboxyl groups. The decrease in O–C
O was coincidental with the decrease of capacitance in gRGO-90 °C-3 h. With further reduction at 5 h, the specific capacitance rose with the increase in carboxyl groups and the enrichment of the π–π* shake-up satellite, which produces additional redox reactions similar to that of O–C
O. Since good electronic transport is required to facilitate the storage of charge at the interfacial sites, gRGOs reduced at 90 °C demonstrated the best specific capacitance due to the enhanced conductivity from the removal of the oxygen functional groups.
Following our previous study on CoC2O4/gRGO composites, we looked at the most suitable gRGO for incorporation with CoC2O4. The composite was synthesized from the reaction between cobalt nitrate precursor with oxalic acid in the gRGO solution.
H2C2O4·2H2O + Co(NO3)2·6H2O + 3H2O → CoC2O4(s) + 11H2O + 2HNO3. | (1) |
The Co loading of the CoC2O4/gRGO from the TGA curves in Fig. S3† agrees well with the computed Co loading from the precursor amounts. For each CoC2O4/gRGO composite, we evaluated the OER catalytic activity using linear sweep voltammetry (LSV) in N2-saturated 0.1 M KOH electrolytes with respect to Hg/HgO as reference electrode. Contrary to the expected result, as shown in the surface plot in Fig. 4, increasing the degree of reduction GO does not guarantee an increase in OER activity. CoC2O4/gRGO composites, where the gRGOs were reduced at 25 °C, showed high OER activities. This behaviour is in stark contrast with the results in Fig. 1, where the gRGOs reduced at 90 °C showed higher Csp. Nevertheless, CoC2O4/gRGO composites reduced at 90 °C exhibited modest OER activity. This unexpected phenomenon has not been reported yet in previous studies using graphene–metal composites for OER.
In our previous study, we have shown that CoC2O4/gRGO demonstrates excellent and stable electrocatalytic activity towards OER, arising from the inherent properties of the components and their physicochemical interaction. As seen in Fig. 5a, CoC2O4/gRGO-25 °C-6 h displayed the best electrocatalytic activity with a current density of 27.9 mA cm−2, which represents 21.1% higher activity than Ir/C catalyst (23.1 mA cm−2) with the same loading amount. In addition, this current density is 70.5% higher than the previous unoptimized CoC2O4/gRGO catalyst in our previous study.24 Evidently, gRGO alone does not provide any catalytic activity, while the OER activity is exclusively due to CoC2O4. The combination of the two components, at the same Co loading, exhibited excellent OER activity. The results in the Tafel slope measurements agree well with that of the LSV. As seen in Fig. S4,† the estimated Tafel slope of CoC2O4/gRGO-25 °C-6 h is the lowest among the Co–graphene composites, and even the Ir/C catalyst. In effect, a lower overpotential is required to increase the current density of OER, leading to a more efficient water splitting reaction.
We believe that the strong interaction between Co and gRGO during the formation of CoC2O4/gRGO is crucial for improving the OER activity. Excellent interaction between the component materials could be attributed to the enriched functional groups in gRGO that are well suited for the sorption of ions and molecules.40–42 The synergistic combination also translates into more durable catalysts, as shown by the long-term electrolysis in Fig. 5b. After 21000 s, Ir black exhibited fast degradation, while the cobalt–graphene composites remained stable. For instance, CoC2O4/gRGO-25 °C-6 h exhibited the best long-term performance with around 42% current reduction after 21
000 s. Even the catalyst prepared at the higher temperature showed better stability than just CoC2O4 and Ir/C. The excellent physical interaction between CoC2O4 and gRGO sheets may have contributed also to the improved stability of the catalyst by preventing further oxidation of cobalt.24
Fig. 6a displays the representative SEM images of CoC2O4/gRGO-25 °C-6 h, showing the homogeneous formation of microstructured CoC2O4 rods with the gRGO. Although various CoC2O4/gRGO catalysts, depending on the reduction condition, catalyze the water oxidation reaction at different rates, their morphologies do not differ much. Furthermore, the TEM image in Fig. 5b displays the wrapping of gRGO sheets on the CoC2O4 structures. Comparing the morphologies of the composites formed under different GO reduction conditions, low temperature reduction allowed the formation of microrods. Without the gRGO, however, cubic-like CoC2O4 structures are formed in the same reduction condition at 25 °C (Fig. S5b†). Furthermore, at 90 °C, an irregularly structured CoC2O4/gRGO was formed (Fig. S5a†). These show how important is the degree of reduction of GO during the nucleation and growth of the CoC2O4 particles.
In order to look at the crystal structure of the composites, we performed X-ray diffraction (XRD) studies. As shown Fig. 6c, the diffractograms of CoC2O4/gRGO catalysts follow the peaks corresponding to CoC2O4 (JCPDS no. 470797). At the same Co loading, it should be noted that the intensity of the (202) diffraction plane at around 19° was higher in CoC2O4/gRGO-25 °C-6 h than CoC2O4/gRGO-90 °C-6 h and CoC2O4 only. This could mean enhanced crystallinity in CoC2O4 in CoC2O4/gRGO-25 °C-6 h that could lead to improved catalytic activity.43 Further, additional minor diffraction peaks between 20° and 55° were observed (JCPDS no. 470797 and 370719), demonstrating the simultaneous formation of CoC2O4 and β-CoC2O4. Although the characteristic GO peak at around 12° disappeared, the graphitic carbon peak at 24° corresponding to RGO did not appear, proving that GO is only partially reduced.
Furthermore, the BET surface area and the pore size distribution (PSD) of gRGOs and CoC2O4/gRGOs were measured. For instance, the specific surface area of gRGO-90 °C-6 h and gRGO-25 °C-6 h were 284 m2 g−1 and 230 m2 g−1, respectively. As shown in Fig. 6d, the N2 adsorption–desorption isotherms and BJH pore size distribution of gRGO-25 °C-6 h and gRGO-90 °C-6 h showed typical isotherm plots of nanoporous structure that are usually found on partially reduced GO.44,45 In contrast, after the combination with CoC2O4, the BET surface areas of CoC2O4/gRGO-25 °C-6 h drastically decreased to 54 m2 g−1, and its isotherm plot approaches to that of pure CoC2O4. This signifies the formation of the CoC2O4 microrods that are in good contact with gRGO sheets. The PSD of CoC2O4/gRGO-25 °C-6 h confirmed the decrease in the pore sizes of the composite. This occurrence, however, was not observed in CoC2O4/gRGO-90 °C-6 h, which had less remaining oxygen functional groups. This supports our assertion on the positive influence of the oxygen functional groups in gRGO during the formation of the composite. It can be said as well that the surface area is not the determining factor in the excellent OER activity and stability of CoC2O4/gRGO-25 °C-6 h, but the synergistic activity between the components.
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c5ra09956k |
‡ These authors contributed equally to this work. |
This journal is © The Royal Society of Chemistry 2015 |