Next Article in Journal
A Comprehensive Study of Blockchain Technology and Its Role in Promoting Sustainability and Circularity across Large-Scale Industry
Previous Article in Journal
Adaptive Grazing of Native Grasslands Provides Ecosystem Services and Reduces Economic Instability for Livestock Systems in the Flooding Pampa, Argentina
Previous Article in Special Issue
Novel Applications of Silk Proteins Based on Their Interactions with Metal Ions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhanced H2 Generation via Piezoelectric Reforming of Waste Sugars and Fruits Using Au-Decorated g-C3N4

1
College of Chemistry and Materials Engineering, Zhejiang A&F University, Hangzhou 311300, China
2
Institute of Advanced Materials and Flexible Electronics (IAMFE), School of Chemistry and Materials Science, Nanjing University of Information Science & Technology, Nanjing 210044, China
3
School of Environmental Science and Engineering, Nanjing University of Information Science & Technology, Nanjing 210044, China
4
Changwang School of Honors, Nanjing University of Information Science & Technology, Nanjing 210044, China
*
Authors to whom correspondence should be addressed.
Sustainability 2024, 16(10), 4231; https://doi.org/10.3390/su16104231
Submission received: 10 April 2024 / Revised: 12 May 2024 / Accepted: 15 May 2024 / Published: 17 May 2024

Abstract

:
The food industry is responsible for generating considerable amounts of waste, such as excess fruits and leftover sugars, which contribute to resource depletion and pose environmental challenges. This research delves into the application of gold-modified graphitic carbon nitride nanosheets (Au/CN) as a potent catalyst for the transformation of these food wastes into H2 via piezoelectric reforming during sonication. Au/CN demonstrated a superior rate of H2 evolution compared to pristine g-C3N4 (i.e., 1533.3 vs. 364.9 µmol/g/h) and it maintained its efficiency through multiple cycles of use. The catalytic activity was found to be optimal at a neutral pH level and with increased sugar concentrations. The enhanced catalytic performance of Au/CN was ascribed to the efficient segregation of charge carriers as well as the reduced charge transfer distance. This study underscores the viability of using Au/CN as a means for converting food wastes into a sustainable source of H2 energy.

1. Introduction

Within the food industry, a substantial volume of waste is produced, most notably in the form of surplus fruits and sugars. These by-products, often dismissed as trivial waste, in fact possess significant potential for reuse and sustainability initiatives. The organic content in these discarded items could be transformed into a variety of useful products, largely through the process of solid-state fermentation [1,2] or thermochemical processes (such as hydrothermal liquefaction [3] and aqueous phase reforming [4]). However, solid-state fermentation, albeit effective, is time-consuming due to its reliance on microbial growth, and it also poses a risk of contamination from unwanted bacteria and fungi. On the other hand, thermochemical methods demand high temperatures and/or pressures. In 2020, the Reisner group unveiled the potential for converting food waste into value-added commodities (predominantly H2) via a photocatalytic reforming process [5]. This new method harnesses photo-generated electrons and holes to drive the reduction of water and the oxidation of food waste, respectively. Requiring no external energy other than sunlight, this technique is capable of producing H2 (from water) that is suitable for fuel cells, thereby offering an enticing fusion of waste treatment and energy production [6,7,8,9]. However, such light-driven reactions are limited to transparent wastewater, as the presence of solids (e.g., fruits) can obstruct light penetration.
Recently, the concept of piezoelectric catalysis, which is driven by mechanical energy, has garnered significant attention [10,11,12]. When piezoelectric materials, lacking a center of symmetry, are subjected to mechanical vibrations, they generate positive and negative charges on the opposite sides of their surface. These charges, reminiscent of the charge separation in photo-reforming reactions, have the potential to catalyze the dissociation of water molecules to release H2 [13], as well as to promote the oxidative degradation of organic compounds [14]. Given that mechanical energy can readily penetrate solutions containing solids or even navigate through opaque solutions without hindrance, piezoelectric catalysis has emerged as a promising approach for food processing applications where light-based methods may be less effective.
The selection and engineering of piezoelectric catalytic materials are critical to the success of H2 production. The present research in piezoelectric catalysis primarily focuses on perovskite-based oxides (such as BaTiO3, PbTiO3, etc. [15,16,17,18,19]), ZnO [20], BiOX [13,21,22] (where X = Cl, Br, I), and MoS2 [23,24]. These materials are not only characterized by high piezoelectric coefficients but are also often structured in one-dimensional (1D) [25] or two-dimensional (2D) [13,26] forms. Such morphological features are conducive to the effective harnessing of mechanical energy through deformation. Nevertheless, these materials require intricate fabrication methods (e.g., hydrothermal and solvent-thermal methods) to attain such specific forms, which presents substantial obstacles to their large-scale production. On the other hand, g-C3N4, known for its unique 2D structure and scalability in synthesis, has emerged as a promising candidate. While predominantly recognized for its role as a photocatalyst [27,28], recent research has unveiled that g-C3N4, featuring triangular nanopores within its tri-s-triazine layers, can display significant in-plane piezoelectric effects. It has been effectively applied in piezoelectric reactions for tasks such as H2 production, H2O2 generation, and the degradation of dyes and antibiotics [29,30,31,32,33]. With its demonstrated efficacy, we believe that g-C3N4 holds great potential as a piezoelectric catalyst to turn waste sugars and fruits into valuable materials by harnessing mechanical energy.
In our current research, we have engineered thin g-C3N4 nanosheets embellished with Au nanoparticles as co-catalysts for H2 production through the piezoelectric reforming of discarded sugars and fruits. The Au/CN composite showcased an enhanced H2 evolution rate in comparison to unmodified g-C3N4. We thoroughly examined the reaction conditions and elucidated the core mechanisms responsible for the observed enhancement in H2 production.

2. Materials and Methods

2.1. Chemicals

Urea (CH4N2O, ≥99.0%), tetrachloroauric (III) acid (HAuCl4, ≥99.9%), chloroplatinic (IV) acid (H2PtCl6, Pt 37.5%), sodium tetrachloropalladate (II) (Na2PdCl4, 98%), silver nitrate (AgNO3, ≥99.0%), and ruthenium chloride hydrate (RuCl3·xH2O, 99.95%) were obtained from Aladdin Holding Group Co., Ltd. (Beijing, China). Sodium sulfate (Na2SO4, ≥99%), D-glucose (C6H12O6, ≥99.5%), fructose (C6H12O6, ≥99%), sucrose (C12H22O11, ≥99.5%), and maltose (C12H22O11, ≥98.0%) were purchased from Sinopharm (Shanghai, China). All these chemicals were used without further purification.

2.2. Preparation

Typically, a crucible containing 10 g of urea was capped and placed in a muffle furnace. The temperature was ramped up at a rate of 5 °C/min to 550 °C and held constant for 4 h. The resultant light-yellow substance was identified as g-C3N4, denoted as CN. For the synthesis of gold-decorated g-C3N4, 10 mg of the as-synthesized CN was dispersed in 20 mL of an aqueous solution with 5 g of glucose and a specified quantity of HAuCl4 to achieve a gold loading of 2 wt.%. The reactor was sealed tightly with a silicone plug and flushed with argon gas through needles for 30 min to expel any air. Subsequently, the mixture was subjected to UV light from a 395 nm LED lamp while stirring magnetically for 1 h. The transformation of the powder from light-yellow to red evidenced the successful deposition of Au nanoparticles onto the g-C3N4, yielding the composite designated as Au2/CN. The sample was collected through centrifugation, cleansed consecutively with water and ethanol, and subsequently oven-dried at 60 °C to ready it for subsequent characterizations. The preparation of Pt-, Pd-, Ag-, and Ru-loaded g-C3N4 used the same method except H2PtCl6, Na2PdCl4, AgNO3, and RuCl3 were used as the precursors, respectively.

2.3. Characterizations

X-ray diffraction (XRD) patterns were obtained using a Shimadzu Lab XRD-6100 X-ray diffractometer (Kyoto, Japan), employing Cu Kα radiation (λ = 0.1541 nm). The AFM (atomic force microscopy) and PFM (Piezoresponse Force Microscopy) analyses were performed utilizing the Bruker Dimension Icon platform (Billerica, MA, USA). X-ray photoelectron spectroscopy (XPS) measurements were conducted on a Shimadzu/Kratos AXIS SUPRA+ instrument (Kyoto, Japan). Transmission electron microscopy (TEM) and high-resolution TEM (HRTEM) images were examined on a Tecnai F20 system operating at an acceleration voltage of 200 kV (Hillsboro, OR, USA). UV–Vis diffuse reflectance spectroscopy (UV-Vis DRS) data were collected via a Shimadzu UV-3101PC spectrophotometer (Kyoto, Japan), utilizing BaSO4 as a standard reference. Photoluminescence (PL) emission spectra were acquired using a Shimadzu RF-6000 spectrometer (Kyoto, Japan), with a 365 nm laser serving as the excitation source.

2.4. Electrochemical Analysis

Electrochemical characterization was performed using a CHI630E electrochemical workstation (Shanghai, China) equipped with a traditional three-electrode configuration, employing a 0.5 M Na2SO4 solution as the electrolyte. A platinum net served as the counter electrode, while an Ag/AgCl electrode saturated with 3.5 M KCl was utilized as the reference electrode, exhibiting a potential of 0.2046 V versus the Normal Hydrogen Electrode (NHE). The working electrode was fashioned by depositing the sample onto fluorine-doped tin oxide (FTO) glass (1 cm × 1 cm) through the following procedure: initially, 10 mg of the sample was dispersed in an 80 μL of Nafion (5%, used as a conductive and stable adhesive) and 5 mL of ethanol mixture using ultrasonic agitation; subsequently, 10 μL of this dispersion was applied onto the surface of the cleaned FTO glass; and ultimately, the coated FTO glass was cured on a hot plate set to 80 °C for a duration of 1 h. Piezoelectric current measurements were conducted by intermittently activating the ultrasonic agitation to acquire the current–time (i-t) response. Electrochemical impedance spectroscopy (EIS) analyses were executed at an applied overpotential of 600 mV over a frequency range from 0.01 Hz to 100 kHz, with a signal amplitude of 5 mV. Mott–Schottky plots were generated within a potential window extending from −1 V to 1 V relative to the Ag/AgCl electrode, at designated frequencies of 0.7, 1.2, and 1.7 kHz.

2.5. Piezoelectric H2 Evolution Test

After the above-mentioned photo-deposition process, the Au2/CN sample immersed in glucose solution (or other sugar solutions) was readied for piezo-catalytic assessment by purging with argon to clear any residual H2 from the process. The glass reactor containing the sample was placed in an ultrasonic cleaner (SN-QX-13, Shanghai, China, operating at 40 kHz, with a 1.3 L capacity and 160 W power), which was filled with water up to a specific level to accentuate vibrations. To ensure consistent test reproducibility, it is crucial to conduct all experiments under the same conditions: utilizing the same ultrasonic cleaner, maintaining the reactor centrally positioned within the sonication bath, and restricting the daily ultrasonication duration to no more than 3 h to circumvent potential attenuation that could result from extended usage. To mitigate thermal effects, the water in the bath was replenished regularly (e.g., every 15–20 min). To quantify the produced H2, 0.1 mL of gas was extracted using a syringe at hourly intervals during sonication and subsequently injected into a gas chromatograph (s·sun GC-9860, Shanghai, China) equipped with a thermal conductivity detector (TCD) for analysis. For tests using fruits as sacrificial reagents, 5 g of fruit pulp was blended with 20 mL of water in a crusher for one minute, ensuring the breakdown of the solid contents. Then, 10 mg of Au2/CN powder was dispersed into the fruit mixture. After an argon purge, the mixture was subjected to the identical sonication and gas sampling protocol.

3. Results and Discussion

g-C3N4 nanosheets were synthesized through the thermal decomposition of urea at 550 °C for a duration of 4 h [27]. The atomic force microscopy (AFM) image presented in Figure 1a illustrates the bidimensional architecture of the material, with lateral dimensions spanning between 1 and 2 µm and a thickness ranging from 4 to 6 nm. To ascertain the ferroelectric attributes of the g-C3N4 nanosheets, piezo-response force microscopy (PFM) was utilized. The amplitude image showcased in Figure 1b displays a pattern of distinct ferroelectric domains scattered randomly, while the phase map in Figure 1c exhibits pronounced phase shifts across various sections of the nanosheets, indicative of heterogeneous polarization directions. In Figure 1d, a local hysteresis loop captures a substantial phase angle shift of approximately 175° upon the application and subsequent reversal of a 10 V DC bias field, suggestive of definitive polarization switching within the g-C3N4 nanosheets [25]. Moreover, the loop’s pronounced butterfly shape in the amplitude profile provides strong confirmation of the material’s piezoelectric response [30].
Au nanoparticles were anchored onto g-C3N4 nanosheets via a photo-deposition technique [27]. As depicted in Figure 2a, the unmodified g-C3N4 demonstrated two distinct XRD peaks at 2θ values of 13.7° and 27.8°, corresponding to the in-plane structural packing and the interlayer stacking of conjugated aromatic systems within the g-C3N4 framework, respectively [27]. Subsequent to Au nanoparticle incorporation, additional peaks at 38.4°, 44.0°, 64.8°, and 77.5° materialized, ascribed, respectively, to the (111), (200), (220), and (311) crystal planes of cubic gold [34]. The ultraviolet–visible diffuse reflectance spectroscopy (UV–Vis DRS) results, shown in Figure 2b, further corroborate the presence of Au by revealing a pronounced absorption peak at approximately 560 nm, which originates from the localized surface plasmon resonance (LSPR) associated with Au nanoparticles [35]. Transmission electron microscopy (TEM) images, both bright and dark field, verified a homogeneous distribution of Au nanoparticles across the g-C3N4 nanosheets, with a narrow size distribution averaging 3.7 ± 1.2 nm. The high-resolution TEM (HRTEM) image inset in Figure 2c exposes a lattice spacing of 0.20 nm, which matches the (200) plane of cubic Au [27], thereby confirming the successful deposition of Au nanoparticles on the g-C3N4 substrate.
Figure 3 presents the XPS spectra of Au2/CN. The survey spectrum identified the characteristic peaks corresponding to C 1s, N 1s, O 1s, and Au 4f core levels. In the detailed C 1s spectrum shown in Figure 3b, three distinct peaks can be observed at 288.3, 286.1, and 284.8 eV, which are attributed to the graphitic N-C=N moieties, C-OH functionalities, and C-C bonding frameworks, respectively [33]. The N 1s spectrum featured in Figure 3c has four peaks at 405.3, 401.6, 400.1, and 399.4 eV, which are, respectively, delineated as π citations [36], N atoms in amine-associated N-H and tertiary nitrogen (N-(C)3), and C-N=C linkages [33]. The Au 4f spectrum had four notable peaks, with the dominant doublet peaks appearing at 84.0 eV and 87.7 eV, which correlate to the 4f5/2 and 4f7/2 orbitals of metallic gold (Au0) [27]. The other doublet peaks at 86.0 eV and 89.7 eV indicate that approximately 25.6% of the Au was in its oxidized state (Au+), which stems from the surface oxidation of gold nanoparticles when exposed to air. This kind of surface oxidation is typical when noble metal particles are diminished to the nanometer scale [27].
The piezoelectric activities of pristine g-C3N4 (CN) and Au2/CN were initially appraised via the sonication of a glucose solution (5 g in 20 mL). As depicted in Figure 4a, H2 production progressed linearly over the course of a three-hour sonication for both samples. However, the H2 yield derived from Au2/CN was markedly superior to that obtained from CN, recording rates of 1533.3 versus 364.9 µmol/g/h, respectively (based on the values noted in Figure 4a). A control experiment, employing a non-piezoelectric material, such as Au/TiO2, as the catalyst resulted in a reaction rate of only 29.6 µmol/g/h, substantiating that H2 was indeed generated through the piezoelectric effect within the g-C3N4 nanosheets. The sonication of pure water without any catalyst revealed a minimal H2 production rate of 13.7 µmol/g/h, effectively dismissing the likelihood of water splitting through sonocatalysis—a chemical reaction propelled by ultrasonic waves [37].
The durability of the Au2/CN sample was evaluated over five testing cycles, showcased in Figure 4b. A significant reduction in H2 production was observed during the second testing cycle, after which, the rate of production stabilized (notably, the fourth cycle exhibited a somewhat reduced production rate, which is attributed to the displacement of the reactor from the central position of the sonication bath). Recent studies have indicated that the sonication process can lead to the exfoliation of g-C3N4, which in turn may result in a diminished piezoelectric response [14]. This phenomenon could account for the observed decrease in H2 evolution during repeated testing. The XRD analysis in Figure 4c revealed a notable decrease in the intensity of the g-C3N4 peak at 27.8° (indicating a decrease in the interlayer stacking of conjugated aromatic systems within the g-C3N4 framework) after the cyclability test, whereas the peak intensity corresponding to Au remained largely unchanged. Furthermore, the TEM image in Figure 4d corroborates the maintained dimensional consistency of the Au nanoparticles and reveals a diminished thickness in the g-C3N4 sheets (compared to Figure 2c).
The subsequent experiments were conducted to elucidate the contribution of metal co-catalysts to the H2 evolution process. Initially, a volcano-shaped trend in activity was noted with ascending Au concentrations, peaking at 2 wt.%. While Au nanoparticles functioned as co-catalysts in the evolution of H2, a superabundance of Au deposited on the g-C3N4 could potentially interfere with the effective capture of mechanical energy. Moving forward, comparisons were made between different metals (2 wt.%), revealing an activity sequence as follows: Au > Ag ≈ Pt ≈ Pd > Ru. Ideally, the inherent H2 evolution activities of these metal nanoparticles should adhere to the order Pt ≈ Pd > Ru > Au > Ag, based on their respective affinities for H* adsorption [38]. To comprehend the significant difference in the observed trend, we performed a TEM analysis on Ag2/CN, Pt2/CN, Pd2/CN, and Ru2/CN. As shown in Figure S1, these four samples exhibited considerable variance in the size of the metal co-catalysts and their dispersion on g-C3N4. For Ag2/CN, scarcely any metal nanoparticles were visible in the TEM mode, indicating that Ag was atomically dispersed on the g-C3N4. The Pt, Pd, and Ru co-catalysts were broadly dispersed on the g-C3N4, with the average size being 1.0 nm, 3.8 nm, and 7.7 nm, respectively. A minor aggregation of nanoparticles was noted in Pd2/CN. Although a concrete conclusion could not been reached yet, the tests suggested that the actual catalytic performance may be substantially affected by other factors such as particle size [39], geometric configuration [40], and the chemical state [41] of the metal components. Our recent investigations also revealed that Au nanoparticles anchored onto g-C3N4 exhibited a larger size and a more pronounced metallic character when compared to Pt nanoparticles synthesized using the identical photo-deposition method, and these Au nanoparticles showcased superior performance in the photocatalytic H2 evolution during the process of glucose reforming [27].
The impact of the solution’s pH and glucose concentration on H2 production was also examined. As illustrated in Figure 5c, the apex of H2 evolution occurred under neutral pH conditions, with a gradual decrease in reaction rate upon lowering the pH. The prevailing theory suggests that at a pH value below the isoelectric point (IEP) of g-C3N4, which is approximately 5 [42], the g-C3N4 assumes a positive charge, thereby diminishing its interaction with protons (H+). It is particularly noteworthy that H2 production was significantly hindered in alkaline solutions (pH = 9–11), contrary to the observations for photocatalysis, where elevated pH values have been reported to improve the H2 evolution by promoting the generation of hydroxyl radicals (OH + h+ → ·OH) [27]. It was proposed that the introduction of NaOH might enhance the solution’s viscosity, thereby dampening the vibrational intensity. Concerning the concentration of glucose, within the range of 0.1 to 1 g per 20 mL, there was no significant impact on the rate of H2 production. However, elevating the glucose concentration to above 5 g per 20 mL resulted in a substantial enhancement of the reaction’s activity.
To elucidate the superior activity of Au2/CN in comparison to its pristine counterpart (CN), the charge separation efficiency of both samples was scrutinized through photoluminescence (PL) spectroscopy and electrochemical analyses. Upon subjecting the samples to 345 nm laser irradiation, PL emission peaks were detected at ~440 nm, as depicted in Figure 6a, which aligns with their adsorption edges shown in Figure 2b. Notably, the more catalytically active Au2/CN demonstrated a diminished PL intensity, indicative of a reduction in positive–negative charge recombination events. Furthermore, when examining the piezoelectric current responses under sonication, Au2/CN manifested a piezo-current approximately twice that of CN, as evidenced in Figure 6b. The EIS spectra in Figure 6c demonstrated a smaller radius of Au2/CN compared to CN, indicating a lower charge-transfer resistance. Collectively, these experiments substantiate the premise that the architectural configuration of Au2/CN is conducive to the enhanced generation and separation of piezoelectrically induced charge carriers. Furthermore, to gain insight into its band edge position, Mott–Schottky analyses of CN were performed. As depicted in Figure 6d, CN demonstrated a flat band potential of −0.65 V versus Ag/AgCl, which corresponds to −0.45 V versus the Normal Hydrogen Electrode (NHE). Taking into account that the conduction band of g-C3N4, being an n-type semiconductor, is typically situated approximately 0.3 V more negative relative to the flat band potential [43], it can be inferred that the conduction band edge of g-C3N4 should lie near −0.75 V versus the NHE. Combined with the 2.90 eV bandgap obtained from the Tauc plot (Figure S2a), the valence band edge was calculated to be 2.15 V versus the NHE, as depicted in the band structure diagram (Figure S2b). This estimation aligns with the values reported in the literature [28] and affirms the capability of CN to facilitate proton reduction to H2 (i.e., more negative than the reduction potential of H2/H+).
The top-performing catalyst, Au2/CN, was put to the test with an assortment of sacrificial reagents, including a range of sugars, fruits, and a beverage (Coca-Cola). The findings, delineated in Figure 7, revealed that out of the four sugars analyzed (namely, glucose, fructose, sucrose, and maltose), glucose demonstrated superior efficacy. This can be attributed to the more hydrolytically amenable structure of glucose’s six-membered pyranose ring, as opposed to the five-membered furanose ring of fructose. Additionally, glucose outstripped the reactivity of disaccharides such as sucrose, which consists of glucose and fructose bound by a glycosidic linkage, and maltose, comprising two glucose units similarly linked. In our assay of various fruits, including apple, watermelon, litchi, banana, grape, orange, and strawberry, we processed 5 g of fruit pulp blended with 20 mL of water in a crusher for one minute to disrupt the solid matter. Subsequently, Au2/CN was introduced directly to the fruit mixture for experimentation. The data depicted in Figure 7 indicate that H2 generation occurred incrementally over time across all fruit samples. Notably, oranges (1568.0 µmol/g/h) and grapes (1385.8 µmol/g/h) exhibited a remarkable H2 production efficiency, which could be ascribed to their higher sugar content, higher glucose ratio, and/or the more fluid nature of their juices. Conversely, banana, in spite of its significant sugar content, resulted in a lower yield of H2 (i.e., 434.7 µmol/g/h). This can be attributed to the dense, viscous consistency of banana pulp, which hampers the seamless diffusion of reactants within the mixture and consequently decelerates the pace of the catalytic H2 production. Furthermore, Coca-Cola, a beverage distinguished by its significant sugar concentration (approximately 2.1 g per 20 mL), underwent evaluation after the elimination of dissolved CO2 via sonication. This led to a somewhat moderate yield of H2, quantified at 107.4 µmol/g/h. This outcome is likely attributed to the predominant use of disaccharides (such as cane sugar) as sweeteners in the Chinese formulation of the drink, and the presence of other additives (e.g., carbonate, caramel color, phosphoric acid, natural flavors, caffeine, etc.) that could potentially inhibit the H2 evolution reaction.
The proposed reaction mechanism is illustrated in Scheme 1. During ultrasound sonication, cavitation bubbles form and crash in the water, exerting transient stress on the CN nanosheets [37]. This stress leads to deformation and lateral stretching of the tri-s-triazine layers, which contain triangular nanopores and subsequently results in in-plane polarization [29]. This polarization induces a charge separation, causing positive and negative charges to migrate towards the peripheries of the CN sheets. In the presence of Au nanoparticles, these charges tend to accumulate on the nanoparticles, which not only promotes a high charge separation efficiency as indicated in reference, but also acts as a co-catalyst for the H2 evolution process [27]. Moreover, given the even distribution of Au nanoparticles across the nanosheets, the charges can be transferred directly to the nanoparticles rather than traveling to the edges of the CN, thereby significantly reducing the charge transfer distance [30]. This shortened path, in conjunction with the catalytic effect of the Au nanoparticles, can markedly amplify the piezoelectric H2 evolution activity. The waste sugars and fruits also served a crucial function within the overall reaction. They acted as scavengers for positive charges to inhibit their recombination with negative charges [5]. Such a role is pivotal in prolonging the lifetime of negative charges, facilitating a more sustained H2 evolution reaction.

4. Conclusions

In summary, g-C3N4-based materials were designed for the conversion of waste sugars and fruits into green fuel H2 through piezoelectric catalysis. The Au2/CN catalyst with the optimal content of Au exhibited remarkable H2 production capabilities, outshining its non-modified counterpart and demonstrating resilience across multiple usage cycles. The study identified the key factors influencing the catalysis process, such as the metal co-catalyst type, pH levels, and sugar concentration. The successful application of Au2/CN with various scavengers for positive charges, and the understanding of charge separation mechanisms, further substantiate its potential for sustainable energy production. This work not only offers a promising avenue for waste-to-energy conversion but also contributes to the broader goal of environmental sustainability by harnessing mechanical energy to stimulate valuable chemical transformations. Nonetheless, the efficiency of H2 production is influenced by a multitude of complex factors that still necessitate further research to optimize and understand the underlying mechanisms.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/su16104231/s1, Figure S1: TEM images of (a) Ag2/CN, (b) Pt2/CN, (c) Pd2/CN and (d) Ru2/CN; Figure S2: (a) The Tauc plot derived from the UV-Vis DRS of Au2/CN, as displayed in Figure 2b. (b) The diagram representing the band structure of g-C3N4.

Author Contributions

Conceptualization, C.W. and F.P.; methodology, C.W.; formal analysis, C.W.; investigation, K.R., F.D. and L.Z.; resources, J.G. and C.W.; data curation, C.W.; writing—original draft preparation, K.R.; writing—review and editing, C.W. and F.P.; supervision, J.G. and C.W.; funding acquisition, J.G. and C.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Natural Science Foundation of Zhejiang Province (LQ22B030009), the National Natural Science Foundation of China (32271528), the Scientific Research Foundation of Zhejiang A and F University (2021FR027), and the Innovation and Entrepreneurship Training Program of ZAFU (S202210341127).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data generated in the study are included in the article. Further inquiries can be directed to the corresponding author.

Acknowledgments

We thank the Shiyanjia Lab (www.shiyanjia.com, accessed on 14 July 2023) for performing the PFM analysis.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Saeed, S.; Shahid, M.; Naseer, R.; Ghazanfar, M.; Irfan, M. Bioconversion of fruit peels to levan by solid state fermentation and statistical optimization by response surface methodology. Biomass Convers. Biorefin. 2023. [Google Scholar] [CrossRef]
  2. Ibarruri, J.; Goiri, I.; Cebrián, M.; García-Rodríguez, A. Solid State Fermentation as a Tool to Stabilize and Improve Nutritive Value of Fruit and Vegetable Discards: Effect on Nutritional Composition, In Vitro Ruminal Fermentation and Organic Matter Digestibility. Animals 2021, 11, 1653. [Google Scholar] [CrossRef] [PubMed]
  3. Demey, H.; Ratel, G.; Lacaze, B.; Delattre, O.; Haarlemmer, G.; Roubaud, A. Hydrogen Production by Catalytic Supercritical Water Gasification of Black Liquor-Based Wastewater. Energies 2023, 16, 3343. [Google Scholar] [CrossRef]
  4. Zoppi, G.; Pipitone, G.; Pirone, R.; Bensaid, S. Aqueous phase reforming process for the valorization of wastewater streams: Application to different industrial scenarios. Catal. Today 2022, 387, 224–236. [Google Scholar] [CrossRef]
  5. Uekert, T.; Dorchies, F.; Pichler, C.M.; Reisner, E. Photoreforming of food waste into value-added products over visible-light-absorbing catalysts. Green Chem. 2020, 22, 3262–3271. [Google Scholar] [CrossRef]
  6. Nagakawa, H.; Nagata, M. Photoreforming of Organic Waste into Hydrogen Using a Thermally Radiative CdOx/CdS/SiC Photocatalyst. ACS Appl. Mater. Interfaces 2021, 13, 47511–47519. [Google Scholar] [CrossRef] [PubMed]
  7. Wu, C.; Fang, L.; Ding, F.; Mao, G.; Huang, X.; Lu, S. Photocatalytic hydrogen production from water and wastepaper on Pt/TiO2 composites. Chem. Phys. Lett. 2023, 826, 140650. [Google Scholar] [CrossRef]
  8. Skillen, N.; Daly, H.; Lan, L.; Aljohani, M.; Murnaghan, C.W.J.; Fan, X.; Hardacre, C.; Sheldrake, G.N.; Robertson, P.K.J. Photocatalytic Reforming of Biomass: What Role Will the Technology Play in Future Energy Systems. Top. Curr. Chem. 2022, 380, 33. [Google Scholar] [CrossRef]
  9. Uekert, T.; Pichler, C.M.; Schubert, T.; Reisner, E. Solar-driven reforming of solid waste for a sustainable future. Nat. Sustain. 2021, 4, 383–391. [Google Scholar] [CrossRef]
  10. Penconi, M.; Rossi, F.; Ortica, F.; Elisei, F.; Gentili, P.L. Hydrogen Production from Water by Photolysis, Sonolysis and Sonophotolysis with Solid Solutions of Rare Earth, Gallium and Indium Oxides as Heterogeneous Catalysts. Sustainability 2015, 7, 9310–9325. [Google Scholar] [CrossRef]
  11. Li, C.; Liu, S.; Zhao, H.; Tian, Y. Performance Assessment and Comparison of Two Piezoelectric Energy Harvesters Developed for Pavement Application: Case Study. Sustainability 2022, 14, 863. [Google Scholar] [CrossRef]
  12. Long, S.X.; Khoo, S.Y.; Ong, Z.C.; Soong, M.F.; Huang, Y.-H. Numerical and Experimental Investigation of a Compressive-Mode Hull Piezoelectric Energy Harvester under Impact Force. Sustainability 2023, 15, 15899. [Google Scholar] [CrossRef]
  13. Long, Y.; Xu, H.; He, J.; Li, C.; Zhu, M. Piezoelectric polarization of BiOCl via capturing mechanical energy for catalytic H2 evolution. Surf. Interfaces 2022, 31, 102056. [Google Scholar] [CrossRef]
  14. He, Q.; Yi, Y.; Shi, W.; Sun, P.; Dong, X. Determination of the key role to affect the piezocatalytic activity of graphitic carbon nitride for tetracycline hydrochloride degradation in water. Chemosphere 2023, 317, 137828. [Google Scholar] [CrossRef] [PubMed]
  15. Charoonsuk, T.; Sriphan, S.; Nawanil, C.; Chanlek, N.; Vittayakorn, W.; Vittayakorn, N. Tetragonal BaTiO3 nanowires: A template-free salt-flux-assisted synthesis and its piezoelectric response based on mechanical energy harvesting. J. Mater. Chem. C 2019, 7, 8277–8286. [Google Scholar] [CrossRef]
  16. Jiang, B.; Iocozzia, J.; Zhao, L.; Zhang, H.; Harn, Y.-W.; Chen, Y.; Lin, Z. Barium titanate at the nanoscale: Controlled synthesis and dielectric and ferroelectric properties. Chem. Soc. Rev. 2019, 48, 1194–1228. [Google Scholar] [CrossRef] [PubMed]
  17. Wang, P.; Li, X.; Fan, S.; Chen, X.; Qin, M.; Long, D.; Tadé, M.O.; Liu, S. Impact of oxygen vacancy occupancy on piezo-catalytic activity of BaTiO3 nanobelt. Appl. Catal. B Environ. 2020, 279, 119340. [Google Scholar] [CrossRef]
  18. Xu, S.; Guo, L.; Sun, Q.; Wang, Z.L. Piezotronic Effect Enhanced Plasmonic Photocatalysis by AuNPs/BaTiO3 Heterostructures. Adv. Funct. Mater. 2019, 29, 1808737. [Google Scholar] [CrossRef]
  19. Zhu, L.; Gu, W.; Zou, W.; Liu, H.; Zhang, Y.; Wu, Q.; Fu, Z.; Lu, Y. Enhancing the Sonolysis Efficiency of SrTiO3 Particles with Cr-Doping. Catal. Lett. 2020, 150, 562–572. [Google Scholar] [CrossRef]
  20. Yuan, J.; Huang, X.; Zhang, L.; Gao, F.; Lei, R.; Jiang, C.; Feng, W.; Liu, P. Tuning piezoelectric field for optimizing the coupling effect of piezo-photocatalysis. Appl. Catal. B Environ. 2020, 278, 119291. [Google Scholar] [CrossRef]
  21. Chen, L.; Zhang, W.; Wang, J.; Li, X.; Li, Y.; Hu, X.; Zhao, L.; Wu, Y.; He, Y. High piezo/photocatalytic efficiency of Ag/Bi5O7I nanocomposite using mechanical and solar energy for N2 fixation and methyl orange degradation. Green Energy Environ. 2021, 8, 283–295. [Google Scholar] [CrossRef]
  22. Li, Z.Q.; Chen, X.H.; Li, T.; Wang, B.J.; Li, N.B.; Luo, H.Q. Crystal facet engineering of polar single crystal BiOCl with improved piezo-photocatalytic activity. Appl. Surf. Sci. 2023, 615, 156283. [Google Scholar] [CrossRef]
  23. Wu, W.; Wang, L.; Li, Y.; Zhang, F.; Lin, L.; Niu, S.; Chenet, D.; Zhang, X.; Hao, Y.; Heinz, T.F.; et al. Piezoelectricity of single-atomic-layer MoS2 for energy conversion and piezotronics. Nature 2014, 514, 470–474. [Google Scholar] [CrossRef] [PubMed]
  24. Lei, R.; Gao, F.; Yuan, J.; Jiang, C.; Fu, X.; Feng, W.; Liu, P. Free layer-dependent piezoelectricity of oxygen-doped MoS2 for the enhanced piezocatalytic hydrogen evolution from pure water. Appl. Surf. Sci. 2022, 576, 151851. [Google Scholar] [CrossRef]
  25. Peng, F.; Lin, J.; Li, H.; Liu, Z.; Su, Q.; Wu, Z.; Xiao, Y.; Yu, H.; Zhang, M.; Wu, C.; et al. Design of piezoelectric ZnO based catalysts for ammonia production from N2 and H2O under ultrasound sonication. Nano Energy 2022, 95, 107020. [Google Scholar] [CrossRef]
  26. Hinchet, R.; Khan, U.; Falconi, C.; Kim, S.-W. Piezoelectric properties in two-dimensional materials: Simulations and experiments. Mater. Today 2018, 21, 611–630. [Google Scholar] [CrossRef]
  27. Ding, F.; Yu, H.; Liu, W.; Zeng, X.; Li, S.; Chen, L.; Li, B.; Guo, J.; Wu, C. Au-Pt heterostructure cocatalysts on g-C3N4 for enhanced H2 evolution from photocatalytic glucose reforming. Mater. Des. 2024, 238, 112678. [Google Scholar] [CrossRef]
  28. Wang, X.; Maeda, K.; Thomas, A.; Takanabe, K.; Xin, G.; Carlsson, J.M.; Domen, K.; Antonietti, M. A metal-free polymeric photocatalyst for hydrogen production from water under visible light. Nat. Mater. 2009, 8, 76–80. [Google Scholar] [CrossRef] [PubMed]
  29. Guo, S.-D.; Mu, W.-Q.; Zhu, Y.-T. Biaxial strain enhanced piezoelectric properties in monolayer g-C3N4. J. Phys. Chem. Solids 2021, 151, 109896. [Google Scholar] [CrossRef]
  30. Hu, C.; Chen, F.; Wang, Y.; Tian, N.; Ma, T.; Zhang, Y.; Huang, H. Exceptional Cocatalyst-Free Photo-Enhanced Piezocatalytic Hydrogen Evolution of Carbon Nitride Nanosheets from Strong In-Plane Polarization. Adv. Mater. 2021, 33, 2101751. [Google Scholar] [CrossRef] [PubMed]
  31. Lei, H.; He, Q.; Wu, M.; Xu, Y.; Sun, P.; Dong, X. Piezoelectric polarization promoted spatial separation of photoexcited electrons and holes in two-dimensional g-C3N4 nanosheets for efficient elimination of chlorophenols. J. Hazard. Mater. 2022, 421, 126696. [Google Scholar] [CrossRef] [PubMed]
  32. Shao, Y.; Liu, C.; Ma, H.; Chen, J.; Dong, C.; Wang, D.; Mao, Z. Piezocatalytic performance difference of graphitic carbon nitride (g-C3N4) derived from different precursors. Chem. Phys. Lett. 2022, 801, 139748. [Google Scholar] [CrossRef]
  33. Vuong, H.-T.; Nguyen, D.-V.; Ly, P.P.; Phan, P.D.M.; Nguyen, T.D.; Tran, D.D.; Mai, P.T.; Hieu, N.H. Defect Engineering of Porous g-C3N4 to Add Multifunctional Groups for Enhanced Production of H2O2 via Piezo-Photocatalysis. ACS Appl. Nano Mater. 2023, 6, 664–676. [Google Scholar] [CrossRef]
  34. Peng, S.; Sun, X.; Sun, L.; Zhang, M.; Zheng, Y.; Su, H.; Qi, C. Selective Hydrogenation of Acetylene Over Gold Nanoparticles Supported on CeO2 Pretreated Under Different Atmospheres. Catal. Lett. 2019, 149, 465–472. [Google Scholar] [CrossRef]
  35. Guo, Z.; Dai, F.; Yin, H.; Zhang, M.; Xing, J.; Wang, L. The dual role of Au nanoparticles in the surface plasmon resonance enhanced photocatalyst Au/g-C3N4. Colloid Interface Sci. Commun. 2022, 48, 100615. [Google Scholar] [CrossRef]
  36. Li, Q.; Wang, Q. Photo(electro)catalyst of Flower-Like Cobalt Oxide Co-Doped g-C3N4: Degradation of Methylene Blue under Visible Light Illumination. Materials 2022, 15, 4104. [Google Scholar] [CrossRef] [PubMed]
  37. Suslick, K.S. Sonochemistry. Science 1990, 247, 1439–1445. [Google Scholar] [CrossRef] [PubMed]
  38. Zeradjanin, A.R.; Vimalanandan, A.; Polymeros, G.; Topalov, A.A.; Mayrhofer, K.J.J.; Rohwerder, M. Balanced work function as a driver for facile hydrogen evolution reaction—Comprehension and experimental assessment of interfacial catalytic descriptor. Phys. Chem. Chem. Phys. 2017, 19, 17019–17027. [Google Scholar] [CrossRef]
  39. Li, X.; Bi, W.; Zhang, L.; Tao, S.; Chu, W.; Zhang, Q.; Luo, Y.; Wu, C.; Xie, Y. Single-Atom Pt as Co-Catalyst for Enhanced Photocatalytic H2 Evolution. Adv. Mater. 2016, 28, 2427–2431. [Google Scholar] [CrossRef] [PubMed]
  40. Cao, S.; Jiang, J.; Zhu, B.; Yu, J. Shape-dependent photocatalytic hydrogen evolution activity over a Pt nanoparticle coupled g-C3N4 photocatalyst. Phys. Chem. Chem. Phys. 2016, 18, 19457–19463. [Google Scholar] [CrossRef] [PubMed]
  41. Wang, Z.; Peng, X.; Tian, S.; Wang, Z. Enhanced hydrogen production from water on Pt/g-C3N4 by room temperature electron reduction. Mater. Res. Bull. 2018, 104, 1–5. [Google Scholar] [CrossRef]
  42. Zhu, B.; Xia, P.; Ho, W.; Yu, J. Isoelectric point and adsorption activity of porous g-C3N4. Appl. Surf. Sci. 2015, 344, 188–195. [Google Scholar] [CrossRef]
  43. Lee, S.F.; Jimenez-Relinque, E.; Martinez, I.; Castellote, M. Effects of Mott–Schottky Frequency Selection and Other Controlling Factors on Flat-Band Potential and Band-Edge Position Determination of TiO2. Catalysts 2023, 13, 1000. [Google Scholar] [CrossRef]
Figure 1. (a) AFM image, (b) amplitude image, (c) phase map, and (d) hysteresis loop of CN.
Figure 1. (a) AFM image, (b) amplitude image, (c) phase map, and (d) hysteresis loop of CN.
Sustainability 16 04231 g001
Figure 2. (a) XRD patterns and (b) UV–Vis DRS of CN and Au2/CN; (c) bright-field TEM image of Au2/CN with an inset showing the HRTEM image of a Au NP; (d) dark field TEM image of Au2/CN with an inset showing the size distribution of the Au nanoparticles.
Figure 2. (a) XRD patterns and (b) UV–Vis DRS of CN and Au2/CN; (c) bright-field TEM image of Au2/CN with an inset showing the HRTEM image of a Au NP; (d) dark field TEM image of Au2/CN with an inset showing the size distribution of the Au nanoparticles.
Sustainability 16 04231 g002
Figure 3. XPS (a) survey spectrum, (b) C 1s spectrum, (c) N 1s spectrum, and (d) Au 4f spectrum of Au2/CN.
Figure 3. XPS (a) survey spectrum, (b) C 1s spectrum, (c) N 1s spectrum, and (d) Au 4f spectrum of Au2/CN.
Sustainability 16 04231 g003
Figure 4. (a) Piezoelectric H2 evolution using CN and Au2/CN catalysts. (b) The recyclability test. (c) The XRD patterns of Au2/CN before and after the recyclability test. (d) The TEM image of Au2/CN after the recyclability test.
Figure 4. (a) Piezoelectric H2 evolution using CN and Au2/CN catalysts. (b) The recyclability test. (c) The XRD patterns of Au2/CN before and after the recyclability test. (d) The TEM image of Au2/CN after the recyclability test.
Sustainability 16 04231 g004
Figure 5. The effect of (a) Au loading, (b) metal type, (c) pH value, and (d) glucose concentration on piezoelectric H2 evolution on Au2/CN.
Figure 5. The effect of (a) Au loading, (b) metal type, (c) pH value, and (d) glucose concentration on piezoelectric H2 evolution on Au2/CN.
Sustainability 16 04231 g005
Figure 6. (a) Photoluminescence, (b) piezo-current response, and (c) EIS of CN and Au2/CN samples. (d) Mott–Schottky plots of CN at frequencies of 0.7, 1.2, and 1.7 kHz.
Figure 6. (a) Photoluminescence, (b) piezo-current response, and (c) EIS of CN and Au2/CN samples. (d) Mott–Schottky plots of CN at frequencies of 0.7, 1.2, and 1.7 kHz.
Sustainability 16 04231 g006
Figure 7. H2 production using different sugars (5 g/20 mL), fruits (5 g/20 mL), and Coca-Cola (20 mL, degassed).
Figure 7. H2 production using different sugars (5 g/20 mL), fruits (5 g/20 mL), and Coca-Cola (20 mL, degassed).
Sustainability 16 04231 g007
Scheme 1. The polarization and charge transfer process on CN and Au2/CN during the piezoelectric H2 evolution reaction.
Scheme 1. The polarization and charge transfer process on CN and Au2/CN during the piezoelectric H2 evolution reaction.
Sustainability 16 04231 sch001
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

Ren, K.; Ding, F.; Zhang, L.; Peng, F.; Guo, J.; Wu, C. Enhanced H2 Generation via Piezoelectric Reforming of Waste Sugars and Fruits Using Au-Decorated g-C3N4. Sustainability 2024, 16, 4231. https://doi.org/10.3390/su16104231

AMA Style

Ren K, Ding F, Zhang L, Peng F, Guo J, Wu C. Enhanced H2 Generation via Piezoelectric Reforming of Waste Sugars and Fruits Using Au-Decorated g-C3N4. Sustainability. 2024; 16(10):4231. https://doi.org/10.3390/su16104231

Chicago/Turabian Style

Ren, Ke, Fangjie Ding, Lijun Zhang, Fengping Peng, Jianzhong Guo, and Chunzheng Wu. 2024. "Enhanced H2 Generation via Piezoelectric Reforming of Waste Sugars and Fruits Using Au-Decorated g-C3N4" Sustainability 16, no. 10: 4231. https://doi.org/10.3390/su16104231

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