Next Article in Journal
Sustainable Microfabrication Enhancement of Graphene Nanoplatelet-Reinforced Biomedical Alumina Ceramic Matrix Nanocomposites
Previous Article in Journal
Solid-Liquid Triboelectric Nanogenerator Based on Vortex-Induced Resonance
Previous Article in Special Issue
Preparation and Characterization of PVDF–TiO2 Mixed-Matrix Membrane with PVP and PEG as Pore-Forming Agents for BSA Rejection
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Insights into the Photoelectrocatalytic Behavior of gCN-Based Anode Materials Supported on Ni Foams

1
Department of Chemical Sciences, Padova University and INSTM, 35131 Padova, Italy
2
Laboratory of Applied Physical and Analytical Chemistry, Department of Inorganic Chemistry, Faculty of Science, University of Yaoundé, Yaoundé P.O. Box 812, Cameroon
3
CNR-ICMATE and INSTM, Department of Chemical Sciences, Padova University, 35131 Padova, Italy
4
EMAT and NANOlab Center of Excellence, University of Antwerp, 2020 Antwerpen, Belgium
5
Section of Chemistry for the Technology (ChemTech), Department of Industrial Engineering, University of Padova and INSTM, 35131 Padova, Italy
6
Interdisciplinary Laboratories for Advanced Materials Physics (i-LAMP), Dipartimento di Matematica e Fisica, Università Cattolica del Sacro Cuore, 25133 Brescia, Italy
*
Authors to whom correspondence should be addressed.
Nanomaterials 2023, 13(6), 1035; https://doi.org/10.3390/nano13061035
Submission received: 16 February 2023 / Revised: 3 March 2023 / Accepted: 6 March 2023 / Published: 13 March 2023
(This article belongs to the Special Issue Advanced Nanocomposite Materials for Water and Wastewater Treatment)

Abstract

:
Graphitic carbon nitride (gCN) is a promising n-type semiconductor widely investigated for photo-assisted water splitting, but less studied for the (photo)electrochemical degradation of aqueous organic pollutants. In these fields, attractive perspectives for advancements are offered by a proper engineering of the material properties, e.g., by depositing gCN onto conductive and porous scaffolds, tailoring its nanoscale morphology, and functionalizing it with suitable cocatalysts. The present study reports on a simple and easily controllable synthesis of gCN flakes on Ni foam substrates by electrophoretic deposition (EPD), and on their eventual decoration with Co-based cocatalysts [CoO, CoFe2O4, cobalt phosphate (CoPi)] via radio frequency (RF)-sputtering or electrodeposition. After examining the influence of processing conditions on the material characteristics, the developed systems are comparatively investigated as (photo)anodes for water splitting and photoelectrocatalysts for the degradation of a recalcitrant water pollutant [potassium hydrogen phthalate (KHP)]. The obtained results highlight that while gCN decoration with Co-based cocatalysts boosts water splitting performances, bare gCN as such is more efficient in KHP abatement, due to the occurrence of a different reaction mechanism. The related insights, provided by a multi-technique characterization, may provide valuable guidelines for the implementation of active nanomaterials in environmental remediation and sustainable solar-to-chemical energy conversion.

1. Introduction

Over the past decades, the ever-increasing industrialization and world population growth have led to an unprecedented consumption of energy, mainly sustained by fossil fuels and, in parallel, to numerous harmful environmental effects, among which is water quality deterioration [1,2,3,4,5,6,7]. The latter is nowadays regarded as a global threat that is negatively impacting natural ecosystems [8] and is precluding large-scale access to clean water resources [9], whose safeguard is an urgent priority towards the wellbeing of society. In fact, the demand for both sustainable energy and drinking quality water is predicted to undergo a notable increase within 2050 [5,10]. Despite the fact that water is a largely abundant natural resource, the most Earth-abundant reservoir (~97% of the total H2O) is salt water from oceans or seas, whereas clean freshwater available for human consumption amounts to less than 1% [11], due to the daily discharge of several harmful compounds [1,2,3,6]. Unfortunately, many of these pollutants (including pesticides, pharmaceuticals, and dyes) are recalcitrant organic species that cannot be efficiently degraded by conventional treatments [1,6,11]. To this aim, a valuable alternative is offered by advanced oxidation processes, enabling efficient H2O purification under ambient conditions thanks to the generation of powerful oxidizing agents (i.e., hydroxyl radicals, •OH) [6,11], as in the Fenton process [10,11,12]. These kinds of photoactivated processes, driven by n-type active semiconductors such as TiO2 [12,13,14,15], are gaining a strategic importance for the simultaneous production of green H2 and water purification, which are of remarkable interest for real-world end-uses towards improved sustainability. As an alternative to TiO2-based photoanodes, gCN, a Vis-light active semiconductor, offers various concomitant advantages, including cost effectiveness, a negligible environmental impact, a favorable chemical reactivity, and good stability in various environments [16]. These features, coupled to the suitable band edge positions, have prompted its increasing investigation as an H2O splitting photoanode [4,5,17,18,19], though its use as a (photo)electrocatalyst for water decontamination has been less explored to date [20].
In the present work, we focused our attention on how relatively inefficient (photo)electrocatalysts for the oxygen evolution reaction (OER), the bottleneck of water splitting, can be, on the contrary, interesting platforms for •OH production in processes aimed at the degradation of aqueous pollutants. Specifically, we show how gCN flakes deposited by EPD on highly conductive and porous Ni foams can efficiently trigger OER if coupled with suitable Co-containing cocatalysts (CoO, CoFe2O4, or CoPi) [21,22]. The latter have been introduced in low amounts onto the pristine gCN by RF-sputtering or electrodeposition. A thorough characterization by complementary analytical tools evidenced that material OER performances were favorably boosted by the formation of suitable nano-junctions, suppressing detrimental electron-hole recombination and the favorable CoPi catalytic action [4,5,19,23]. Conversely, the use of bare gCN resulted in a more efficient production of •OH radicals and an enhanced degradation of aqueous KHP, a recalcitrant pollutant used as a standard benchmark [24,25] as confirmed by the “coumarin test” [14,26]. This aspect, i.e., the enhanced •OH radicals production over supported gCN flakes, which has not been addressed in detail up to date, is a key issue for the eventual mineralization of water soluble organic pollutants. In the following, after a multi-technique chemico-physical investigation of the developed electrocatalyst materials, their functional performances are presented and critically discussed as a function of the used synthesis and processing conditions. This work provides a new paradigm for constructing advanced photocatalysts for effective organic pollutant degradation and green energy conversion.

2. Materials and Methods

2.1. Material Preparation

In the present work, gCN-based systems were fabricated either in an Ar or air atmosphere (hereafter indicated as gCNAr or gCNair, respectively). Specifically, gCN powders (synthesized as described in the Supplementary Materials Section S1.1) were finely grinded and subsequently used in EPD experiments. Accordingly, 40 mg of gCN and 10 mg of I2 were dispersed into a beaker containing 50 mL of acetone and the resulting mixture was sonicated for 20 min. Then, a pre-cleaned [27] Ni foam substrate (cathode) and carbon paper (anode), connected to a DC generator, were immersed in the resulting suspension at a distance of 10 mm. For bare and functionalized gCNAr-based materials, experimental parameters for the carbon nitride deposition (EPD conditions: 5 min/50 V) were optimized according to our recently published work [19]. For gCNair-electrode materials, carbon nitride EPD conditions were preliminarily adjusted, testing different deposition times (from 1 to 15 min) and applied potentials (from 10 to 50 V), and selecting the ones (1 min/10 V) yielding the best photocurrents. Finally, the samples obtained from gCNAr and gCNair powders were thermally treated in Ar (550 °C, 2.5 h) or in air (300 °C, 1 h), respectively.
The functionalization of gCNAr-based samples with CoO or CoFe2O4 cocatalysts was performed by RF-sputtering from Co (Alfa Aesar®, Ward Hill, MA, USA; purity = 99.95%) or Co3O4-Fe2O3 targets (Neyco, Vanves, France; purity = 99.9%), using a custom-built, two-electrode plasmochemical reactor (ν = 13.56 MHz) and with the previously reported experimental settings [19]. The resulting materials were ultimately annealed in an Ar atmosphere at 500 °C for 2.5 h. Conversely, gCNair-based electrodes were functionalized with CoPi by electrodeposition. Experiments were carried out using an electrochemical workstation (Autolab PGSTAT-204 potentiostat/galvanostat, Utrecht, The Netherlands) and a three-electrode system composed of a Ni foam-supported material as the working electrode, a Pt coil as the counter-electrode, and an Ag/AgCl reference electrode. CoPi deposition was carried out from 0.5 mM CoCl2 aqueous solutions in 0.1 M potassium phosphate buffer (PBS, i.e., 0.1 M K2HPO4 and KH2PO4, pH = 7.1), following a previously reported procedure [23], until the highest OER current density was achieved (typically within 2–3 CV cycles). Figure S1 reports the overall preparation procedure for gCN powders and the resulting Ni foam electrode materials, along with the most relevant synthesis conditions.

2.2. Material Characterization

2.2.1. Characterization of gCN Powders

X-ray diffraction (XRD) measurements were performed using a Bruker (Karlsruhe, Germany) AXS D8 Advance Plus diffractometer, equipped a CuKα X-ray source (λ = 1.54051 Å). The average crystal size was estimated through the Scherrer equation. Analyses were carried out at the PanLab facility (Department of Chemical Sciences, Padova University) founded by the MIUR Dipartimento di Eccellenza grant “NExuS”. Fourier transform-infrared (FT-IR) spectra were recorded in transmittance mode on KBr pellets by means of a Thermo-Nicolet (Nicolet Instrument Corporation, WI, USA) Nexus 860 instrument (resolution = 4 cm−1). UV-Vis diffuse reflectance spectra were recorded on a Cary 5E (Varian, Palo Alto, CA, USA) spectrophotometer (spectral bandwidth = 1 nm), equipped with an integration sphere. The evaluation of band gap (EG) values was performed using the Tauc equation [f(R)]n vs. , where f(R) is the Kubelka–Munk function and R is the measured reflectance, assuming indirect and allowed transitions (n = ½) [18,28,29].

2.2.2. Characterization of Electrode Materials

X-ray photoelectron spectroscopy (XPS) analyses were performed on a multiscan system (Omicron, ScientaOmicron, Uppsala, Sweden) using a Mg X-ray source (1253.6 eV) and a Phoibos 100 SPECS (SPECS, Berlin, Germany) analyzer. The binding energy values were corrected for charging phenomena by assigning a position of 284.8 eV to the adventitious C1s component [30]. After a Shirley-type background subtraction, curve fitting was carried out by the XPSpeak (Version 4.1) software [31], using Gaussian-Lorentzian sum functions. Atomic percentages (at.%) were calculated by peak area integration. Field emission scanning electron microscopy (FE-SEM) analyses were performed by collecting in-lens and backscattered electron signals using a Zeiss (Oberkochen, Germany) SUPRA 40VP instrument at primary beam acceleration voltages of 10–20 kV. Energy dispersive X-ray spectroscopy (EDXS) maps were acquired with an INCA x-act PentaFET Precision spectrometer. The analysis of electrode materials by transmission electron microscopy (TEM) required sample preparation using a ThermoFisher Scientific (Waltham, MA, USA) Helios Nanolab 650 dual-beam focused ion beam instrument. A Pt protection layer was deposited in order to avoid material damage, and the obtained sample foils were thinned to a thickness < 50 nm. Electron diffraction, EDXS, high-resolution scanning TEM (HRSTEM), electron energy loss spectroscopy (EELS), and image simulation were employed for a detailed characterization. The electron diffraction and EDXS analyses were performed on a ThermoFisher Scientific Osiris microscope equipped with a Super-X windowless EDX detector system, operated at 200 kV. High-resolution EELS spectra were acquired using a state-of-the-art double-corrected and monochromated ThermoFisher Scientific Titan 80-300 microscope operated at 120 kV. Further details are reported in the Supplementary Materials (Section S1.2).

2.3. Functional Tests

OER electrochemical tests were carried out both in the dark and under irradiation, using an Autolab three-electrode workstation (PGSTAT204 potentiostat/galvanostat), with the prepared samples and a Pt coil serving as a working electrode and counter-electrode. An Hg/HgO electrode, typically employed in alkaline media, was selected as the reference. The electrolyte was a 0.1 M KOH (pH = 12.9) aqueous solution. The working electrode was exposed to a white light LED source (intensity ≈ 150 mW/cm2). For pollutant degradation, the cell was filled with 50 mL of aqueous solution containing 1 mM KHP and 0.1 M KOH. Then, Ni foam-supported samples were tested as (photo)anodes under constant stirring at a fixed bias voltage of 1.5 V vs. the reversible hydrogen electrode (RHE). Hereafter, all potentials are expressed with respect to RHE. At regular time intervals, aliquots of the KHP solution were collected and the residual pollutant content was measured by flow injection analysis combined with electrospray mass spectrometry (FIA-ESI/MS) [32]. For each sample, FIA-ESI/MS measurements were carried out three times, revealing a <3% deviation among repeated experiments. FIA-ESI/MS analyses were performed with a LCQFleet ion trap instrument (ThermoFisher Scientific), operating in negative ion mode, coupled with a Surveyor LC Pump Plus (ThermoFisher Scientific). The entrance capillary temperature, ion source temperature, and voltage were set to 275 °C, 300 °C, and 4 kV, respectively. The initial KHP solutions were diluted 10 times with H2O/acetonitrile (50/50, v/v) with 0.1% of formic acid. The resulting solutions (25 µL) were injected into the mass spectrometer, using a mixture of H2O/acetonitrile (80/20, v/v) with 0.1% formic acid as eluent (flow rate = 0.3 µL/min). To ensure reproducibility and reliability of the obtained results, each degradation experiment was repeated twice.
The “coumarin test” was used to probe •OH radical production by monitoring the formation of photoluminescent 7-hydroxy-coumarin [14,26,33]. Although such method does not allow for an absolute quantitation of •OH radicals, it is a highly sensitive and valuable tool for the determination of relative activities between different photocatalysts [33]. To this aim, electrochemical experiments were carried out under illumination on a 1 mM coumarin solution in 0.1 M PBS, using the above-described cell and experimental conditions. Fluorescence spectra were collected in the 342–700 nm wavelength range using an FLS 1000 fluorimeter (Edinburgh Instruments, Livingston, United Kingdom). The following settings were used: excitation wavelength/bandwidth = 332/1.5 nm; emission bandwidth = 2.5 nm; optical path = 1 cm. Measurements were also carried out on Ni foam-supported materials using the same instrument, with an excitation wavelength of 410 nm.

3. Results and Discussion

As already mentioned, gCN powders were synthesized by the thermal condensation of melamine at 550 °C for 4 h either in Ar (gCNAr) or in air (gCNair), to investigate the influence of processing conditions on the properties and performances of the resulting Ni foam-supported samples.
The XRD patterns of the gCN powders synthesized in Ar and air are compared in Figure 1a. In both cases, the reflections at ≈13.2° and ≈27.3° were attributed to the packing of tri-s-triazine units in the (100) crystallographic plane, and to the interplanar (002) stacking of gCN sheets, respectively [17,34,35]. The broad and weak peaks at ≈44° and ≈57° can be ascribed to gCN (300) and (004) planes [35]. The corresponding mean nanocrystal sizes were directly affected by the used reaction atmosphere (Figure 1a, inset).
The FT-IR spectra of gCNAr and gCNair powders (Figure 1b) displayed similar features, with a multi-component band in the region of 1200–1650 cm−1 due to typical stretching modes of aromatic heptazine (C6N7) heterocycles, along with the characteristic heptazine breathing mode at 808 cm−1 [17,34,36]. The broad band in the 2900–3500 cm−1 range resulted from stretching vibrations of the –OH groups at the edges of carbon nitride sheets [36,37] and of uncondensed –NHx groups (x = 1, 2), whose presence was also responsible for the N–H deformation mode at 885 cm−1 [17,18,19,36]. The presence of such moieties was confirmed by XPS data (see below). The analysis of gCN powders by UV-Vis diffuse reflectance spectroscopy evidenced a relatively steep edge at λ ≈ 500 nm (Figure S2a), due to interband electronic transitions [38,39]. Tauc plot analyses yielded EG values of ≈2.6 eV for both gCNair and gCNAr (Figure S2b) [17,19].
Both kinds of powders were subsequently used as precursors for EPD on the Ni foams of gCN systems, which were eventually functionalized with CoO and CoFe2O4 (via RF-sputtering), or CoPi (via electrodeposition). Figure 2 reports representative SEM images for samples gCNair and gCNair–CoPi, along with EDXS elemental maps for the latter sample. As can be observed, µm-sized gCN particles with a flake-like morphology were uniformly dispersed over the Ni foam, yet leaving uncovered some substrate regions. For gCNair–CoPi, EDXS maps for carbon and nitrogen highlighted the discontinuous deposit nature. Nonetheless, the O and P signals overlapped with the gCN ones, thus indicating that CoPi deposition preferentially took place over carbon nitride grains rather than on the uncovered Ni foam regions. Although cobalt could not be revealed by EDXS, its presence was clearly confirmed by XPS analyses (see below and the Supplementary Materials).
As far as gCNAr-containing samples are concerned, a lower Ni foam coverage was revealed (Figure S3). In the case of functionalized specimens, the presence of ultra-dispersed CoO or CoFe2O4 nanoparticles (samples gCNAr–CoO and gCNAr–CoFe2O4, respectively) was clearly revealed by high resolution TEM analyses. More specifically, the former specimen contained spherical aggregates with a size smaller than 10 nm. The d-spacings and high angle annular dark-field high resolution scanning TEM (HAADF-HRSTEM) results (Figure 3a) were in line with the occurrence of cubic CoO. This conclusion was confirmed not only by the matching between experimental and simulated HRSTEM images (Figure 3a), but even by the electronic structure analysis of the target nanoparticles. EELS measurements revealed indeed a good agreement between the experimental Co L2,3 (Figure 3b) and O K (Figure S4a) edges and those of a CoO reference, thus allowing for the exclusion of Co3O4 presence. For specimen gCNAr–CoFe2O4, TEM analysis on the observed nanoparticles highlighted the formation of spinel-type CoFe2O4 aggregates (Figure 3c). Even in this case, a good agreement was observed between experimental and simulated HRSTEM images, and the formation of phase-pure CoFe2O4 was confirmed by the EELS spectra of Co L2,3, Fe L2,3, and O K edges (Figure 3d and Supplementary Figure S4b).
The surface chemical composition and elemental chemical states were investigated by XPS. Survey spectra (Figure S5) displayed the presence of C and N photopeaks and, for gCNAr–CoO and gCNAr–CoFe2O4 specimens, even the presence of Co or Co and Fe signals. The occurrence of Ni peaks indicated an incomplete coverage of the underlying Ni foam. The quantitative analyses (see data in the caption for Figure S5) suggested the occurrence of nitrogen-deficient systems. As observed in Figure 4a, Supplementary Figures S6 and S7a, the C1s photopeak could be fitted by three components (see also Figure 4b): a, due to adventitious contamination (BE = 284.8 eV) [40,41]; b, related to C atoms in uncondensed C–NHx (x = 1, 2) groups on carbon nitride ring edges (BE = 286.2 eV) [42,43]; c, due to C in N-C=N moieties (BE = 288.2 eV) [28,40,43,44]. The weight of component b underwent an appreciable increase upon going from bare gCNAr, to gCNAr–CoO, up to gCNAr–CoFe2O4 (Figure 4c). These data suggested an enhancement of defects related to the presence of amino groups, resulting from gCNAr bombardment during RF-sputtering processes, according to the order gCNAr < gCNAr–CoO < gCNAr–CoFe2O4. Recent reports have demonstrated that the presence of surface defects in electrode materials can effectively tune the electronic structure and charge density, and thus improve the redox reactivity [45,46]. In particular, the above-mentioned N defects can beneficially suppress charge carriers recombination, boosting the ultimate material photoactivity [42,47] (see also below). These conclusions were in line with the N1s peak fitting results (see Figures S7b and S8 and related comments).
For both gCNAr–CoO and gCNAr–CoFe2O4 samples, the Co2p photopeak (Figure 4d) presented well-developed shake-up structures on the high BE side of the main components, indicating the occurrence of Co(II) centers [28,30,48]. In fact, for gCN–CoO, the BE positions [BE(Co2p3/2) = 781.1 eV; spin-orbit separation (SOS) = 15.7 eV] were in line with previously reported data for CoO [28,29]. In the case of gCNAr–CoFe2O4, Co2p3/2 BE was slightly higher (781.3 eV), in agreement with the data pertaining to CoFe2O4 systems [36,49]. The formation of the latter oxide, confirmed by TEM and related analyses (see above), was also supported by the Fe2p peak features (Figure 4e; BE(Fe2p3/2) = 710.6 eV; SOS = 13.3 eV) [36,49,50]. In the case of gCNair–CoPi, the Co2p spectral features were in accordance with the literature data for CoPi systems (Figure S7c) [51]. O1s spectra are reported in the Supplementary Materials (see Figure S9 and pertaining observations).
Figure 5a,b display linear sweep voltammetry scans collected on the target specimens, where the current increase at potentials > 1.5 V vs. RHE corresponds to the OER. As can be observed, for Ar-treated samples (Figure 5a) the difference between light and dark currents was relatively small (0.1–0.4 mA/cm2 at E > 1.6 V). This result was related to a high and uneven carbon nitride loading onto Ni foams, negatively affecting the system conductivity and producing a modest light response. This phenomenon, in turn, could be ascribed to the potential delay caused by the resistance–capacitance couple in series, particularly evident at high scan rates [52]. In a different way, gCNair and gCNair–CoPi, containing a lower gCN amount (Figure 5b), yielded much higher photocurrent densities (1.0–2.5 mA/cm2 at E > 1.6 V), due to a more uniform gCN flakes distribution.
It is worth highlighting that both Ar- and air-treated samples revealed a remarkable current density improvement upon gCN functionalization, with either CoO, CoFe2O4, or CoPi. The anodic peaks between 1.35 and 1.5 V correspond to the MO → MO(OH) reaction usually occurring on the surface of Ni and Co oxide nanoparticles [27]. The higher peak intensity for NiFAr and gCNAr in comparison to the homologous gCNair and NiFair samples could be attributed to an enhanced Ni oxidation, related, in turn, to a less uniform substrate coverage by gCN in the case of Ar-treated samples. The resulting systems were tested as (photo)electrocatalysts for the purification of aqueous solutions containing KHP, a recalcitrant pollutant, whose decomposition is obtained in the presence of •OH radicals [54]. Preliminary degradation tests were carried out to assess the electrochemical performances of gCNAr–CoO and gCNAr–CoFe2O4 both in the dark and under irradiation (Figure S10). As can be observed, the specimens yielded a ~50% decrease in the KHP concentration after 1 h, while longer times did not result in further significant variations. Such a result suggests that the decrease in KHP concentration vs. time was mainly due to KHP adsorption on gCN flakes via π–π interactions [55]. A slight additional decrease for longer times, especially under illumination, was observed for gCNAr–CoFe2O4. This result was also in line with indications provided by the “coumarin test” (see below, Figure 7a), revealing that only a modest •OH production took place in the case of gCNAr–CoO, whereas for gCNAr–CoFe2O4, •OH formation was enhanced.
Based on the higher gCNair and gCNair–CoPi photoresponses (Figure 5), such specimens were subsequently tested in light-activated KHP degradation, analyzing the process kinetics by FIA-ESI/MS. Whereas the sample gCNair promoted a significant and continuous decrease in the KHP concentration as a function of time, gCNair–CoPi was almost completely inactive (Figure 6a). Interestingly, the electrochemical performances of gCNair towards KHP degradation were preserved upon repeating the same experiment after three months, highlighting a good material stability. The experimental trend in Figure 6b, in agreement with a pseudo-first order kinetics [56], could be fitted with either a single or a double exponential function. The obtainment of a better match with experimental results in the latter case (see the corresponding χ2 value in Figure 6b) indicates that two distinct processes, i.e., adsorption and decomposition, take place simultaneously on the material surface.
The different behaviors of gCNair and gCNair–CoPi, apparently unexpected since an opposite activity trend was detected in the OER (Figure 5), can be explained considering that organic pollutant degradation is directly dependent on •OH production and does not involve the four-electron mechanism of OER. In alkaline media, the OER process involves in fact the adsorption of hydroxyl ions on the catalyst active sites. Subsequently, hydroxyl ions are oxidized to hydroxyl radicals that are hence deprotonated and oxidized again, yielding surface-adsorbed oxygens. If two of such species are close enough, they can finally react together to produce an O2 molecule [57]. In a different way, KHP degradation is mostly dependent on the anodic production of •OH radicals, whose formation takes place on the catalyst surface when electrogenerated holes react with adsorbed H2O/OH species [2,57]. The attack of •OH radicals to KHP typically involves the initial hydroxylation of the aromatic ring and/or decarboxylation of its side chains with the formation of several hydroxybenzoic acids and benzene diols. Then, the aromatic ring is opened and the formation of several low molecular weight carboxylic acids takes place, followed by further oxidation steps, finally yielding to CO2 and H2O [54,58,59].
To attain a deeper insight into the above processes, photoluminescence experiments with coumarin solutions were carried out to probe the formation of the luminescent 7-hydroxy-coumarin, occurring in the presence of hydroxyl radicals [14,26]. Such tests clearly revealed a significant production of •OH radicals only for the sample gCNair, whereas a negligible photoluminescence was detected for gCNair–CoPi (Figure 7a), in good agreement with the relative KHP degradation trends of Figure 6a. How should this result be interpreted?
As reported by Ehrmaier et al. [60], the production of •OH radicals on the gCN surface under illumination occurs on heptazine ring units. Heptazine photoexcitation to the lowest π–π* singlet yields the formation of an electron-hole pair (Figure 7b). The generated hole can subsequently oxidize a water molecule attached to heptazine via a hydrogen bond. The subsequent electron transfer from water to heptazine is followed by the release of the hydrogen-bonded proton, resulting in the final formation of heptazinyl and •OH radicals. This mechanism can experience appreciable variations in the presence of surface trap states for holes (Figure 7b, lower part), as Co(II) centers in the Co-containing cocatalysts, whose occurrence appreciably reduces the effective hydroxyl radical content. We believe that photogenerated holes are trapped on Co centers so that the mechanism occurring on CoPi nanoparticles is activated, while the oxidation of OH ions on heptazine units is depressed. In fact, solid-state luminescence spectra collected on the samples gCNair and gCNair–CoPi (Figure 8) revealed that CoPi introduction completely modified the charge transport dynamics on the gCN surface, producing a strong luminescence quenching attributable to Co(II) trap states.

4. Conclusions

In this study, gCN flakes were grown on Ni foam substrates by EPD under different conditions, and eventually functionalized with low amounts of highly dispersed CoO, CoFe2O4, or CoPi cocatalysts. Preliminary water splitting experiments revealed that OER performances of the developed systems were improved by gCN functionalization with the above Co-containing species. In a different way, when the same materials were tested as photoelectrocatalysts for KHP degradation, the most active system turned out to be bare gCN, free from any cocatalyst. This apparently surprising behavior was traced back to the different reaction mechanisms involved in water splitting or KHP degradation. In particular, for bare gCN as such, the enhanced generation of •OH radicals (very limited, on the contrary, for systems functionalized with Co-based cocatalysts) efficiently triggered the target pollutant degradation, as demonstrated by means of complementary characterization studies.
The applicative potential of the present work is highlighted by the fact that these electrodes can be easily prepared even with much larger dimensions, and can be easily recovered for subsequent recycling at variance with powdered gCN photocatalysts. In perspective, the outcomes of our research activities yield a promising prospect for the design and engineering of active nanomaterials featuring tailored properties for solar-to-chemical energy conversion and environmental remediation.

Supplementary Materials

The following are available online at https://www.mdpi.com/2079-4991/13/6/1035/s1, optical spectra; additional experimental details and FE-SEM, EELS, XPS, and photodegradation data.

Author Contributions

Investigation, S.B., M.B., C.M., A.G., R.S., N.G., A.O., G.P., V.B., V.D.N. and L.G.; data curation, J.V., D.B., A.G. and G.A.R.; supervision, C.M. and G.A.R.; writing—original draft preparation, A.G. and G.A.R.; writing—review and editing, S.B., M.B., C.M. and D.B.; project administration, D.B. and C.M.; funding acquisition, D.B., C.M. and V.D.N. All authors have read and agreed to the published version of the manuscript.

Funding

The present work was financially supported by CNR (Progetti di Ricerca @CNR—avviso 2020—ASSIST), Padova University (P-DiSC#04BIRD2020-UNIPD EUREKA, DOR 2020–2022), AMGA Foundation (NYMPHEA project), INSTM Consortium (INSTM21PDGASPAROTTO—NANOMAT, INSTM21PDBARMAC—ATENA) and the European Union’s Horizon 2020 research and innovation program under grant agreement No. 823717—ESTEEM3. The FWO-Hercules fund G0H4316N ‘Direct electron detector for soft matter TEM’ is also acknowledged. Many thanks are also due to Dr. Riccardo Lorenzin for his support to experimental activities.

Data Availability Statement

Data supporting this study are available within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Pang, X.; Skillen, N.; Gunaratne, N.; Rooney, D.W.; Robertson, P.K.J. Removal of phthalates from aqueous solution by semiconductor photocatalysis: A review. J. Hazard. Mater. 2021, 402, 123461. [Google Scholar] [CrossRef] [PubMed]
  2. Wang, G.; Zhang, Q.; Chen, Q.; Ma, X.; Xin, Y.; Zhu, X.; Ma, D.; Cui, C.; Zhang, J.; Xiao, Z. Photocatalytic degradation performance and mechanism of dibutyl phthalate by graphene/TiO2 nanotube array photoelectrodes. Chem. Eng. J. 2019, 358, 1083–1090. [Google Scholar] [CrossRef]
  3. Singh, N.; Goldsmith, B.R. Role of electrocatalysis in the remediation of water pollutants. ACS Catal. 2020, 10, 3365–3371. [Google Scholar] [CrossRef]
  4. Ren, Y.; Zeng, D.; Ong, W.-J. Interfacial engineering of graphitic carbon nitride (g-C3N4)-based metal sulfide heterojunction photocatalysts for energy conversion: A review. Chin. J. Catal. 2019, 40, 289–319. [Google Scholar] [CrossRef]
  5. Zou, X.; Sun, Z.; Hu, Y.H. g-C3N4-based photoelectrodes for photoelectrochemical water splitting: A review. J. Mater. Chem. A 2020, 8, 21474–21502. [Google Scholar] [CrossRef]
  6. Garcia-Segura, S.; Brillas, E. Applied photoelectrocatalysis on the degradation of organic pollutants in wastewaters. J. Photochem. Photobiol. C 2017, 31, 1–35. [Google Scholar] [CrossRef]
  7. Kandel, D.R.; Kim, H.-J.; Lim, J.-M.; Poudel, M.B.; Cho, M.; Kim, H.-W.; Oh, B.-T.; Nah, C.; Lee, S.H.; Dahal, B.; et al. Cold plasma-assisted regeneration of biochar for dye adsorption. Chemosphere 2022, 309, 136638. [Google Scholar] [CrossRef]
  8. Poudel, M.B.; Awasthi, G.P.; Kim, H.J. Novel insight into the adsorption of Cr(VI) and Pb(II) ions by MOF derived Co-Al layered double hydroxide @hematite nanorods on 3D porous carbon nanofiber network. Chem. Eng. J. 2021, 417, 129312. [Google Scholar] [CrossRef]
  9. Guo, H.; Niu, H.-Y.; Wang, W.-J.; Wu, Y.; Xiong, T.; Chen, Y.-R.; Su, C.-Q.; Niu, C.-G. Schottky barrier height mediated Ti3C2 MXene based heterostructure for rapid photocatalytic water disinfection: Antibacterial efficiency and reaction mechanism. Sep. Purif. Technol. 2023, 312, 123412. [Google Scholar] [CrossRef]
  10. Brillas, E.; Martínez-Huitle, C.A. Decontamination of wastewaters containing synthetic organic dyes by electrochemical methods. An updated review. Appl. Catal. B 2015, 166–167, 603–643. [Google Scholar] [CrossRef]
  11. Brillas, E.; Sirés, I.; Oturan, M.A. Electro-Fenton process and related electrochemical technologies based on Fenton’s reaction chemistry. Chem. Rev. 2009, 109, 6570–6631. [Google Scholar] [CrossRef] [PubMed]
  12. Brillas, E.; Boye, B.; Sirés, I.; Garrido, J.A.; Rodríguez, R.M.A.; Arias, C.; Cabot, P.-L.S.; Comninellis, C. Electrochemical destruction of chlorophenoxy herbicides by anodic oxidation and electro-Fenton using a boron-doped diamond electrode. Electrochim. Acta 2004, 49, 4487–4496. [Google Scholar] [CrossRef]
  13. Flox, C.; Cabot, P.-L.; Centellas, F.; Garrido, J.A.; Rodríguez, R.M.; Arias, C.; Brillas, E. Solar photoelectro-Fenton degradation of cresols using a flow reactor with a boron-doped diamond anode. Appl. Catal. B 2007, 75, 17–28. [Google Scholar] [CrossRef]
  14. Nakabayashi, Y.; Nosaka, Y. OH radical formation at distinct faces of rutile TiO2 crystal in the procedure of photoelectrochemical water oxidation. J. Phys. Chem. C 2013, 117, 23832–23839. [Google Scholar] [CrossRef]
  15. Poudel, M.B.; Yu, C.; Kim, H.J. Synthesis of conducting bifunctional polyaniline@Mn-TiO2 nanocomposites for supercapacitor electrode and visible light driven photocatalysis. Catalysts 2020, 10, 546. [Google Scholar] [CrossRef]
  16. Wang, W.; Niu, Q.; Zeng, G.; Zhang, C.; Huang, D.; Shao, B.; Zhou, C.; Yang, Y.; Liu, Y.; Guo, H.; et al. 1D porous tubular g-C3N4 capture black phosphorus quantum dots as 1D/0D metal-free photocatalysts for oxytetracycline hydrochloride degradation and hexavalent chromium reduction. Appl. Catal. B 2020, 273, 119051. [Google Scholar] [CrossRef]
  17. Kessler, F.K.; Zheng, Y.; Schwarz, D.; Merschjann, C.; Schnick, W.; Wang, X.; Bojdys, M.J. Functional carbon nitride materials—Design strategies for electrochemical devices. Nature Rev. Mater. 2017, 2, 17030. [Google Scholar] [CrossRef]
  18. Benedet, M.; Rizzi, G.A.; Gasparotto, A.; Lebedev, O.I.; Girardi, L.; Maccato, C.; Barreca, D. Tailoring oxygen evolution performances of carbon nitride systems fabricated by electrophoresis through Ag and Au plasma functionalization. Chem. Eng. J. 2022, 448, 137645. [Google Scholar] [CrossRef]
  19. Benedet, M.; Rizzi, G.A.; Gasparotto, A.; Gauquelin, N.; Orekhov, A.; Verbeeck, J.; Maccato, C.; Barreca, D. Functionalization of carbon nitride systems by cobalt and cobalt-iron oxides boosts solar water oxidation performances. Appl. Surf. Sci. 2023, 618, 156652. [Google Scholar] [CrossRef]
  20. Liang, F.; Wang, H.; Yu, R.; Liu, C.; Wang, Y.; Bai, L.; Hao, C.; Hao, G. Recent progress in photoelectrocatalysis of g-C3N4 for water environment remediation. Prog. Nat. Sci. Mater. Int. 2022, 32, 538–553. [Google Scholar] [CrossRef]
  21. Khosravi, M.; Mohammadi, M.R. Trends and progress in application of cobalt-based materials in catalytic, electrocatalytic, photocatalytic, and photoelectrocatalytic water splitting. Photosynth. Res. 2022, 154, 329–352. [Google Scholar] [CrossRef] [PubMed]
  22. Li, X.; Niu, Z.; Jiang, J.; Ai, L. Cobalt nanoparticles embedded in porous N-rich carbon as an efficient bifunctional electrocatalyst for water splitting. J. Mater. Chem. A 2016, 4, 3204–3209. [Google Scholar] [CrossRef]
  23. Kanan, M.W.; Nocera, D.G. In situ formation of an oxygen-evolving catalyst in neutral water containing phosphate and Co2+. Science 2008, 321, 1072–1075. [Google Scholar] [CrossRef] [PubMed]
  24. Aktaş, Y.; Gözmen, B.; Oturan, M.A. Degradation of phthalic acid by anodic oxidation in acidic aqueous solutions with high chromium content using boron-doped diamond anode. Sep. Purif. Technol. 2022, 293, 121098. [Google Scholar] [CrossRef]
  25. Wang, P.; Ding, Y.; Zhu, L.; Zhang, Y.; Zhou, S.; Xie, L.; Li, A. Oxidative degradation/mineralization of dimethyl phthalate (DMP) from plastic industrial wastewater using ferrate(VI)/TiO2 under ultraviolet irradiation. Environ. Sci. Pollut. Res. 2022, 29, 15159–15171. [Google Scholar] [CrossRef]
  26. Nakabayashi, Y.; Nosaka, Y. The pH dependence of OH radical formation in photo-electrochemical water oxidation with rutile TiO2 single crystals. Phys. Chem. Chem. Phys. 2015, 17, 30570–30576. [Google Scholar] [CrossRef] [PubMed]
  27. Maccato, C.; Bigiani, L.; Girardi, L.; Gasparotto, A.; Lebedev, O.I.; Modin, E.; Barreca, D.; Rizzi, G.A. Plasma-assisted synthesis of Co3O4-based electrocatalysts on Ni foam substrates for the oxygen evolution reaction. Adv. Mater. Interfaces 2021, 8, 2100763. [Google Scholar] [CrossRef]
  28. Zhang, D.; Peng, L.; Liu, K.; Garcia, H.; Sun, C.; Dong, L. Cobalt nanoparticle with tunable size supported on nitrogen-deficient graphitic carbon nitride for efficient visible light driven H2 evolution reaction. Chem. Eng. J. 2020, 381, 122576. [Google Scholar] [CrossRef]
  29. Ismael, M.; Wark, M. Photocatalytic activity of CoFe2O4/g-C3N4 nanocomposite toward degradation of different organic pollutants and their inactivity toward hydrogen production: The role of the conduction band position. FlatChem 2022, 32, 100337. [Google Scholar] [CrossRef]
  30. Briggs, D.; Seah, M.P. Practical Surface Analysis: Auger and X-ray Photoelectron Spectroscopy, 2nd ed.; Wiley: New York, NY, USA, 1990. [Google Scholar]
  31. Available online: https://xpspeak.software.informer.com/4.1/ (accessed on 20 February 2020).
  32. Ruzicka, J.; Hansen, E.H. Flow Injection Analysis, 2nd ed.; Springer: New York, NY, USA, 1988. [Google Scholar]
  33. Žerjav, G.; Albreht, A.; Vovk, I.; Pintar, A. Revisiting terephthalic acid and coumarin as probes for photoluminescent determination of hydroxyl radical formation rate in heterogeneous photocatalysis. Appl. Catal. A 2020, 598, 117566. [Google Scholar] [CrossRef]
  34. Dong, F.; Li, Y.; Wang, Z.; Ho, W.-K. Enhanced visible light photocatalytic activity and oxidation ability of porous graphene-like g-C3N4 nanosheets via thermal exfoliation. Appl. Surf. Sci. 2015, 358, 393–403. [Google Scholar] [CrossRef]
  35. Fina, F.; Callear, S.K.; Carins, G.M.; Irvine, J.T.S. Structural investigation of graphitic carbon nitride via XRD and neutron diffraction. Chem. Mater. 2015, 27, 2612–2618. [Google Scholar] [CrossRef]
  36. Görmez, Ö.; Yakar, E.; Gözmen, B.; Kayan, B.; Khataee, A. CoFe2O4 nanoparticles decorated onto graphene oxide and graphitic carbon nitride layers as a separable catalyst for ultrasound-assisted photocatalytic degradation of Bisphenol-A. Chemosphere 2022, 288, 132663. [Google Scholar] [CrossRef] [PubMed]
  37. Zhou, L.; Lei, J.; Wang, F.; Wang, L.; Hoffmann, M.R.; Liu, Y.; In, S.-I.; Zhang, J. Carbon nitride nanotubes with in situ grafted hydroxyl groups for highly efficient spontaneous H2O2 production. Appl. Catal. B 2021, 288, 119993. [Google Scholar] [CrossRef]
  38. Yu, Y.; Wu, S.; Gu, J.; Liu, R.; Wang, Z.; Chen, H.; Jiang, F. Visible-light photocatalytic degradation of bisphenol A using cobalt-to-oxygen doped graphitic carbon nitride with nitrogen vacancies via metal-to-ligand charge transfer. J. Hazard. Mater. 2020, 384, 121247. [Google Scholar] [CrossRef]
  39. Zhang, G.; Zang, S.; Wang, X. Layered Co(OH)2 deposited polymeric carbon nitrides for photocatalytic water oxidation. ACS Catal. 2015, 5, 941–947. [Google Scholar] [CrossRef]
  40. Wang, J.; Zhang, W.-D. Modification of TiO2 nanorod arrays by graphite-like C3N4 with high visible light photoelectrochemical activity. Electrochim. Acta 2012, 71, 10–16. [Google Scholar] [CrossRef]
  41. Sima, M.; Vasile, E.; Sima, A.; Preda, N.; Logofatu, C. Graphitic carbon nitride based photoanodes prepared by spray coating method. Int. J. Hydrog. Energy 2019, 44, 24430–24440. [Google Scholar] [CrossRef]
  42. Yu, H.; Shi, R.; Zhao, Y.; Bian, T.; Zhao, Y.; Zhou, C.; Waterhouse, G.I.N.; Wu, L.-Z.; Tung, C.-H.; Zhang, T. Alkali-assisted synthesis of nitrogen deficient graphitic carbon nitride with tunable band structures for efficient visible-light-driven hydrogen evolution. Adv. Mater. 2017, 29, 1605148. [Google Scholar] [CrossRef]
  43. Nasri, A.; Jaleh, B.; Nezafat, Z.; Nasrollahzadeh, M.; Azizian, S.; Jang, H.W.; Shokouhimehr, M. Fabrication of g-C3N4/Au nanocomposite using laser ablation and its application as an effective catalyst in the reduction of organic pollutants in water. Ceram. Int. 2021, 47, 3565–3572. [Google Scholar] [CrossRef]
  44. Ren, Y.; Feng, D.; Yan, Z.; Sun, Z.; Zhang, Z.; Xu, D.; Qiao, C.; Chen, Z.; Jia, Y.; Chan Jun, S.; et al. Interfacial coupled engineering of plasmonic amorphous MoO3-x nanodots/g-C3N4 nanosheets for photocatalytic water splitting and photothermal conversion. Chem. Eng. J. 2023, 453, 139875. [Google Scholar] [CrossRef]
  45. Kang, L.; Zhang, M.; Zhang, J.; Liu, S.; Zhang, N.; Yao, W.; Ye, Y.; Luo, C.; Gong, Z.; Wang, C.; et al. Dual-defect surface engineering of bimetallic sulfide nanotubes towards flexible asymmetric solid-state supercapacitors. J. Mater. Chem. A 2020, 8, 24053–24064. [Google Scholar] [CrossRef]
  46. Liu, S.; Kang, L.; Hu, J.; Jung, E.; Henzie, J.; Alowasheeir, A.; Zhang, J.; Miao, L.; Yamauchi, Y.; Jun, S.C. Realizing superior redox kinetics of hollow bimetallic sulfide nanoarchitectures by defect-induced manipulation toward flexible solid-state supercapacitors. Small 2022, 18, 2104507. [Google Scholar] [CrossRef]
  47. Xiao, Y.; Tian, G.; Li, W.; Xie, Y.; Jiang, B.; Tian, C.; Zhao, D.; Fu, H. Molecule self-assembly synthesis of porous few-layer carbon nitride for highly efficient photoredox catalysis. J. Am. Chem. Soc. 2019, 141, 2508–2515. [Google Scholar] [CrossRef] [PubMed]
  48. Chagas, C.A.; de Souza, E.F.; de Carvalho, M.C.N.A.; Martins, R.L.; Schmal, M. Cobalt ferrite nanoparticles for the preferential oxidation of CO. Appl. Catal. A 2016, 519, 139–145. [Google Scholar] [CrossRef]
  49. Wang, G.; Ma, Y.; Wei, Z.; Qi, M. Development of multifunctional cobalt ferrite/graphene oxide nanocomposites for magnetic resonance imaging and controlled drug delivery. Chem. Eng. J. 2016, 289, 150–160. [Google Scholar] [CrossRef]
  50. Palanivel, B.; Lallimathi, M.; Arjunkumar, B.; Shkir, M.; Alshahrani, T.; Al-Namshah, K.S.; Hamdy, M.S.; Shanavas, S.; Venkatachalam, M.; Ramalingam, G. rGO supported g-C3N4/CoFe2O4 heterojunction: Visible-light-active photocatalyst for effective utilization of H2O2 to organic pollutant degradation and OH radicals production. J. Environ. Chem. Eng. 2021, 9, 104698. [Google Scholar] [CrossRef]
  51. Zhang, R.; van Straaten, G.; di Palma, V.; Zafeiropoulos, G.; van de Sanden, M.C.M.; Kessels, W.M.M.; Tsampas, M.N.; Creatore, M. Electrochemical activation of atomic layer-deposited cobalt phosphate electrocatalysts for water oxidation. ACS Catal. 2021, 11, 2774–2785. [Google Scholar] [CrossRef] [PubMed]
  52. Anantharaj, S.; Noda, S.; Driess, M.; Menezes, P.W. The pitfalls of using potentiodynamic polarization curves for Tafel analysis in electrocatalytic water splitting. ACS Energy Lett. 2021, 6, 1607–1611. [Google Scholar] [CrossRef]
  53. Niu, P.; Yin, L.-C.; Yang, Y.-Q.; Liu, G.; Cheng, H.-M. Increasing the visible light absorption of graphitic carbon nitride (Melon) photocatalysts by homogeneous self-modification with nitrogen vacancies. Adv. Mater. 2014, 26, 8046–8052. [Google Scholar] [CrossRef]
  54. Garcia-Segura, S.; Salazar, R.; Brillas, E. Mineralization of phthalic acid by solar photoelectro-Fenton with a stirred boron-doped diamond/air-diffusion tank reactor: Influence of Fe3+ and Cu2+ catalysts and identification of oxidation products. Electrochim. Acta 2013, 113, 609–619. [Google Scholar] [CrossRef]
  55. Liu, J.; Shi, H.; Shen, Q.; Guo, C.; Zhao, G. A biomimetic photoelectrocatalyst of Co–porphyrin combined with a g-C3N4 nanosheet based on π–π supramolecular interaction for high-efficiency CO2 reduction in water medium. Green Chem. 2017, 19, 5900–5910. [Google Scholar] [CrossRef]
  56. Tian, M.; Wu, G.; Adams, B.; Wen, J.; Chen, A. Kinetics of photoelectrocatalytic degradation of nitrophenols on nanostructured TiO2 electrodes. J. Phys. Chem. C 2008, 112, 825–831. [Google Scholar] [CrossRef]
  57. Dau, H.; Limberg, C.; Reier, T.; Risch, M.; Roggan, S.; Strasser, P. The mechanism of water oxidation: From electrolysis via homogeneous to biological catalysis. ChemCatChem 2010, 2, 724–761. [Google Scholar] [CrossRef]
  58. Deng, D.; Wu, X.; Li, M.; Qian, S.; Tang, B.; Wei, S.; Zhang, J. Electrochemical degradation of three phthalate esters in synthetic wastewater by using a Ce-doped Ti/PbO2 electrode. Chemosphere 2020, 259, 127488. [Google Scholar] [CrossRef] [PubMed]
  59. Souza, F.L.; Aquino, J.M.; Irikura, K.; Miwa, D.W.; Rodrigo, M.A.; Motheo, A.J. Electrochemical degradation of the dimethyl phthalate ester on a fluoride-doped Ti/β-PbO2 anode. Chemosphere 2014, 109, 187–194. [Google Scholar] [CrossRef]
  60. Ehrmaier, J.; Domcke, W.; Opalka, D. Mechanism of photocatalytic water oxidation by graphitic carbon nitride. J. Phys. Chem. Lett. 2018, 9, 4695–4699. [Google Scholar] [CrossRef]
Figure 1. (a) XRD patterns of gCNAr and gCNair powders. The average gCN crystallite sizes along the [100] and [002] directions are reported in the inset. (b) Corresponding FT-IR spectra.
Figure 1. (a) XRD patterns of gCNAr and gCNair powders. The average gCN crystallite sizes along the [100] and [002] directions are reported in the inset. (b) Corresponding FT-IR spectra.
Nanomaterials 13 01035 g001
Figure 2. FE-SEM micrographs for samples (a,b) gCNair and (c,d) gCNair–CoPi. In panel (a), left and right images were obtained collecting back-scattered (BSE) and secondary (SE) electrons, respectively. The gCN presence corresponds to dark regions in (a). EDXS elemental maps for specimen gCNair–CoPi recorded on the electron image in (e) are also reported.
Figure 2. FE-SEM micrographs for samples (a,b) gCNair and (c,d) gCNair–CoPi. In panel (a), left and right images were obtained collecting back-scattered (BSE) and secondary (SE) electrons, respectively. The gCN presence corresponds to dark regions in (a). EDXS elemental maps for specimen gCNair–CoPi recorded on the electron image in (e) are also reported.
Nanomaterials 13 01035 g002
Figure 3. HAADF-HRSTEM images (left) and EELS spectra (right) of the Co L2,3 and eventually Fe L2,3 edges acquired on Co-containing nanoparticles for samples gCNAr–CoO (a,b) and gCNAr–CoFe2O4 (c,d). ‘S’ marks the simulated HRSTEM images.
Figure 3. HAADF-HRSTEM images (left) and EELS spectra (right) of the Co L2,3 and eventually Fe L2,3 edges acquired on Co-containing nanoparticles for samples gCNAr–CoO (a,b) and gCNAr–CoFe2O4 (c,d). ‘S’ marks the simulated HRSTEM images.
Nanomaterials 13 01035 g003
Figure 4. XPS characterization of gCNAr deposits on Ni foams before and after functionalization with CoO and CoFe2O4 via RF-sputtering. (a) C1s photoelectron peaks for gCNAr–CoO and gCNAr–CoFe2O4. (b) Sketch of gCN structure [53], in which non-equivalent C sites are marked. Color codes as in panel (a). (c) Percentage contribution of the b component to the overall C1s photopeak for the target specimens. Calculation was performed excluding the adventitious component a. (d) Co2p and (e) Fe2p signals for gCNAr–CoO and gCNAr–CoFe2O4. In (d), stars (*) indicate shake-up peaks.
Figure 4. XPS characterization of gCNAr deposits on Ni foams before and after functionalization with CoO and CoFe2O4 via RF-sputtering. (a) C1s photoelectron peaks for gCNAr–CoO and gCNAr–CoFe2O4. (b) Sketch of gCN structure [53], in which non-equivalent C sites are marked. Color codes as in panel (a). (c) Percentage contribution of the b component to the overall C1s photopeak for the target specimens. Calculation was performed excluding the adventitious component a. (d) Co2p and (e) Fe2p signals for gCNAr–CoO and gCNAr–CoFe2O4. In (d), stars (*) indicate shake-up peaks.
Nanomaterials 13 01035 g004
Figure 5. Linear sweep voltammetry scans collected in 0.1 M KOH solution on (a) Ar-treated and (b) air-treated supported samples. Continuous and dotted curves have been collected under irradiation and in the dark, respectively. For each sample set, bare Ni foams (NiF) scans are also reported.
Figure 5. Linear sweep voltammetry scans collected in 0.1 M KOH solution on (a) Ar-treated and (b) air-treated supported samples. Continuous and dotted curves have been collected under irradiation and in the dark, respectively. For each sample set, bare Ni foams (NiF) scans are also reported.
Nanomaterials 13 01035 g005
Figure 6. (a) KHP degradation tests carried out on gCNair and gCNair–CoPi. KHP residual content was determined from the m/z = 165 hydrogen phthalate peak area as follows: AreaSX/AreaSM × 100, where SM = mother solution; S0 = solution after adsorption/desorption equilibrium in the absence of external bias; SX = solution after testing each specimen under irradiation for X hours (X = 1÷5). For sample gCNair, experiment 2 (red curve) was performed three months after experiment 1 (green plot). (b) Average of the two plots reported in (a) for sample gCNair and fit of the corresponding experimental results both with a single or double exponential function.
Figure 6. (a) KHP degradation tests carried out on gCNair and gCNair–CoPi. KHP residual content was determined from the m/z = 165 hydrogen phthalate peak area as follows: AreaSX/AreaSM × 100, where SM = mother solution; S0 = solution after adsorption/desorption equilibrium in the absence of external bias; SX = solution after testing each specimen under irradiation for X hours (X = 1÷5). For sample gCNair, experiment 2 (red curve) was performed three months after experiment 1 (green plot). (b) Average of the two plots reported in (a) for sample gCNair and fit of the corresponding experimental results both with a single or double exponential function.
Nanomaterials 13 01035 g006
Figure 7. (a) Luminescence spectra obtained from 1 mM coumarin solution in 0.1 M phosphate buffer after 1, 2, or 3 h of photoelectrochemical work under illumination. Curves corresponding to bare Ni foam are also displayed. The luminescence peak at 390 nm is due to coumarin emission [14]. The increased emission at ~450 nm for gCNair indicates an appreciably higher production of 7-hydroxy-coumarin for this specimen. (b) Sketch of the mechanism for the production of •OH radicals on gCN surface [60]. The upper and lower parts of the image sketch are processes taking place in the absence and in the presence of trap states [Co(II) centers], respectively.
Figure 7. (a) Luminescence spectra obtained from 1 mM coumarin solution in 0.1 M phosphate buffer after 1, 2, or 3 h of photoelectrochemical work under illumination. Curves corresponding to bare Ni foam are also displayed. The luminescence peak at 390 nm is due to coumarin emission [14]. The increased emission at ~450 nm for gCNair indicates an appreciably higher production of 7-hydroxy-coumarin for this specimen. (b) Sketch of the mechanism for the production of •OH radicals on gCN surface [60]. The upper and lower parts of the image sketch are processes taking place in the absence and in the presence of trap states [Co(II) centers], respectively.
Nanomaterials 13 01035 g007
Figure 8. Solid-state luminescence spectra collected on samples gCNair and gCNair–CoPi. The spectrum for the bare Ni foam is also reported for comparison.
Figure 8. Solid-state luminescence spectra collected on samples gCNair and gCNair–CoPi. The spectrum for the bare Ni foam is also reported for comparison.
Nanomaterials 13 01035 g008
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

Benedoue, S.; Benedet, M.; Gasparotto, A.; Gauquelin, N.; Orekhov, A.; Verbeeck, J.; Seraglia, R.; Pagot, G.; Rizzi, G.A.; Balzano, V.; et al. Insights into the Photoelectrocatalytic Behavior of gCN-Based Anode Materials Supported on Ni Foams. Nanomaterials 2023, 13, 1035. https://doi.org/10.3390/nano13061035

AMA Style

Benedoue S, Benedet M, Gasparotto A, Gauquelin N, Orekhov A, Verbeeck J, Seraglia R, Pagot G, Rizzi GA, Balzano V, et al. Insights into the Photoelectrocatalytic Behavior of gCN-Based Anode Materials Supported on Ni Foams. Nanomaterials. 2023; 13(6):1035. https://doi.org/10.3390/nano13061035

Chicago/Turabian Style

Benedoue, Serge, Mattia Benedet, Alberto Gasparotto, Nicolas Gauquelin, Andrey Orekhov, Johan Verbeeck, Roberta Seraglia, Gioele Pagot, Gian Andrea Rizzi, Vincenzo Balzano, and et al. 2023. "Insights into the Photoelectrocatalytic Behavior of gCN-Based Anode Materials Supported on Ni Foams" Nanomaterials 13, no. 6: 1035. https://doi.org/10.3390/nano13061035

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