Next Article in Journal
Pt-Based Intermetallic Nanocrystals in Cathode Catalysts for Proton Exchange Membrane Fuel Cells: From Precise Synthesis to Oxygen Reduction Reaction Strategy
Next Article in Special Issue
Synthesis, Property Characterization and Photocatalytic Activity of the Ag3PO4/Gd2BiTaO7 Heterojunction Catalyst under Visible Light Irradiation
Previous Article in Journal
The Catalysis Effect of Na and Point Defect on NO Heterogeneous Adsorption on Carbon during High-Sodium Zhundong Coal Reburning: Structures, Interactions and Thermodynamic Characteristics
Previous Article in Special Issue
Sustainable Hydrogen Production from Starch Aqueous Suspensions over a Cd0.7Zn0.3S-Based Photocatalyst
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Biochar Nanoparticles over TiO2 Nanotube Arrays: A Green Co-Catalyst to Boost the Photocatalytic Degradation of Organic Pollutants

1
Department of Science and High Technology, University of Insubria, Via Valleggio 11, 22100 Como, Italy
2
Photo-Catalytic Synthesis Group, MESA+ Institute for Nanotechnology, University of Twente, P.O. Box 217, 7500 AE Enschede, The Netherlands
3
National Research Council of Italy (CNR), Istituto SCITEC, Via C. Golgi 19, 20133 Milan, Italy
4
Department of Theoretical and Applied Sciences, University of Insubria, Via Dunant, 3, 22100 Varese, Italy
*
Authors to whom correspondence should be addressed.
Submission received: 5 August 2021 / Revised: 20 August 2021 / Accepted: 29 August 2021 / Published: 29 August 2021

Abstract

:
Biochar nanoparticles (BC NPs), produced by low temperature pyrolysis (350 °C) of microalgae (Nannochloropsis sp.) and nutshells, are proposed as low-cost and sustainable co-catalysts to promote the photocatalytic activity of TiO2 nanotube (NT) arrays towards the degradation of methylene blue (MB) used as an organic pollutant model molecule. BC NPs (size < 25 nm) were obtained by treating bulk BC (i.e., biomass after pyrolysis) by sonication–centrifugation cycles in a water solution. The filtered BC NPs dispersion was deposited by simple drop-casting on the TiO2 NT support. The BC loading was varied by performing multiple depositions. Photocatalytic experiments under UV light (365 nm) revealed that the decoration with BC NPs significantly improves the TiO2 photoactivity. Such enhancement is mainly influenced by the amount of BC deposited; upon optimizing the BC deposition conditions, the rate of photocatalytic degradation of MB increases approximately three times with respect to bare TiO2, almost irrespective of the nature of the raw material. The greater photocatalytic activity of BC-TiO2 can be attributed to the synergistic combination of reactant/product adsorption and catalytic degradation of the adsorbed organic pollutant, as well as an improved charge carrier separation and electron transfer.

Graphical Abstract

1. Introduction

With the accelerated expansion of worldwide industrialization and urbanization, environmental pollution is increasingly becoming a serious global issue faced by humankind [1]. Therefore, green, cost-effective, and efficient ways to resolve this adverse situation have been extensively studied by researchers in recent decades. Among the various strategies conceived to solve this global problem, heterogeneous photocatalysis, in which the naturally available and clean energy of the sun can be exploited by semiconductors (e.g., TiO2, α-Fe2O3, ZnO), has emerged as one of the most promising technologies, especially for water purification [2,3]. Photocatalytic materials have been widely studied for the degradation and removal of many inorganic [4,5,6] and organic [7,8] pollutants in natural and waste waters.
Although a broad variety of semiconductors have been explored so far, titanium dioxide (TiO2) is still the most studied material, owing to its non-toxicity, stability against (photo)corrosion, and low cost [9]. Nanostructured TiO2 (e.g., nanoparticles, nanotubes, nanorods) is extensively employed as it can provide a larger surface area and more efficient charge collection (transfer) by redox species in the reaction environment compared to bulk TiO2 [10]. In addition, co-catalysts forming heterojunctions on TiO2 are also employed to further promote charge separation and transfer [11,12,13] or to enhance the catalytic activity (e.g., providing new reaction sites and pathways or increasing reactant adsorption) [6,13,14]; metal oxides (e.g., WO3, Fe2O3) [15,16,17], metals (e.g., Au, Pt) [11,18], or metal alloys [19,20] are typically used for this purpose. Carbon-based materials (e.g., carbon nanotubes [21,22], graphene oxide [23], and carbon dots [24,25]) have also been proposed as alternative co-catalysts or supports for TiO2-based photocatalysts.
In this context, biochar (BC), a porous carbonaceous material produced from biomass feedstocks via pyrolysis, has recently been proposed as a greener supporting carbon material for powdered photocatalytic platforms [26,27,28] and, in a fewer reports, as co-catalytic nanoparticles supported on nanostructured TiO2 [29]. As one main advantage, the employment of BC allows for strongly reducing the environmental footprint and costs associated with the photocatalyst synthesis compared to traditional noble metal- and carbon-based materials; its production is highly sustainable, as waste biomasses can be used as raw materials [30,31]. Many other advantages of using biochar are reported in the literature concerning the implementation of renewable energies as easy metal recovery [32] and its application in fuel cells [33]. The unique physicochemical properties of BC make it an ideal candidate for (photo)catalytic applications due to its high specific surface area, easily tunable functional groups, chemical stability, and electrical conductivity [27,34,35]. The coupling of BC with TiO2 can therefore offer a synergistic combination of reactant/product adsorption and catalytic degradation of the adsorbed pollutants, as well as an improved charge carrier separation and electron transfer system, which result in a significantly enhanced overall photocatalytic efficiency [27,36].
The properties of BC can be tuned by changing pyrolysis temperature and biomass feedstocks [37,38,39] in order to reach the best (photo)catalytic performances. Highly conductive, graphitized BC produced at high temperatures (800–1000 °C), characterized by a high specific surface area and porosity, is often preferred for catalytic applications, as BC is commonly employed as a support for photocatalyst nanoparticles [27,40]. Activating agents (e.g., KOH, H3PO4) are also employed to further promote the creation of pores and to enable the formation of surface functional groups, which are partially lost at high temperatures due to the graphitization of the material [41]. At temperatures above 500–600 °C, in fact, the degradation of organic matter with stronger chemical bonds occurs, leading to a loss of functionalization. Despite their high performances, the required harsh pyrolysis conditions drastically reduce (i) the environmental friendliness of BC fabrication, (ii) BC yield [37], and (iii) the natural self-doping of BC (which strictly depends on the feedstock chemical composition and pyrolysis temperature) [42]. In contrast, the employment of milder conditions, e.g., chemical-free slow pyrolysis at low temperatures (<400 °C), can provide a solution to these limitations leading to high production yields of naturally functionalized and self-doped BC with improved adsorptive properties [43,44,45,46]. Considering that functionalized materials are highly sought in (photo)catalytic applications, the combination of low temperature pyrolysis with heteroatom-rich biomass feedstocks, e.g., microalgae, should be beneficial, as highly functionalized biochar with improved adsorption properties is produced. However, the limited specific surface area of BC produced at low temperature hinders its typical employment as supporting material for powdered photocatalysts, and that is why it is more rarely studied. Downsizing of BC to quasi-spherical nanoparticles, e.g., by forming carbon dots by sonication, centrifugation, and filtration [47]), may instead open the possibility to exploit such highly functionalized BC nanoparticles as co-catalytic material deposited onto nanostructured photocatalytic layers; the BC downsizing process results in a greatly increased specific surface area and enhanced electronic and optical properties [47,48]. Moreover, the production of BC NPs should be not considered less green than bulk materials, as the employment of chemicals or harsh conditions (that may hinder the sustainability of the production process) is not involved. In this light, BC NPs can be deemed as a green co-catalyst that can enable the possibility to fabricate a supported photocatalytic layer rather than conventional powdered photocatalytic platforms produced by using bulk biochar.
In light of these considerations, BC nanoparticles produced by the slow pyrolysis at a low temperature (350 °C) of microalgae (Nannochloropsis sp.) and nutshell were used in this work as a co-catalyst on the surface of anodic TiO2 nanotube arrays (TiO2 NTs) to boost the photocatalytic activity towards the degradation of methylene blue, selected as an organic pollutant model. These biomass feedstocks were selected because of their markedly different composition, with microalgae-derived BC being noticeably more polar than the nutshell-derived one. The investigation of these diverse materials aims at evaluating the effect of the nature of the starting biomass feedstock on the promotion of the catalytic performances. The influence of other parameters such as BC loading and photocatalytic conditions were evaluated, too.

2. Results and Discussion

2.1. Bulk BC Characterization

Microalgae- and nutshell-derived biochar used for the fabrication of BC-TiO2 photocatalysts were produced and fully characterized in a previous work [46]. Briefly, BC having high surface rugosity (see SEM images in Figure S1) was produced with quite high yield by using both lignocellulosic and microalgae feedstocks (34 ± 2% and 41 ± 4% yields, respectively). Concerning the elemental composition of these materials, heteroatom content (P and N) of MBC is tenfold of that observed for NBC (Figure S2): this evidence is coherent with the higher amount of protein and lipids in microalgae. Furthermore, a more marked presence of aromatic C–H, as well as of –OH, C=O, C–O–C, and N-H functional groups, was found in MBC compared to NBC, in which no carbonyl bands were detected. Such features can be ascribed to the diverse nature of the feedstocks and to the catalytic effect of alkali (Na and K) and alkaline earth (Mg and Ca) metals during pyrolysis (see Figure S2 for alkali and alkaline earth contents); these chemicals, massively present in microalgae, can promote the formation of light oxygenates via ring fragmentation reactions of carbohydrates [46]. Finally, XRD analysis showed that no crystalline phases were formed during pyrolysis, most likely due to the low temperature of the thermal treatment [41].

2.2. Characterization of TiO2 and BC-TiO2 Photocatalysts

After pyrolysis, bulk BC powder was dispersed in ultrapure water (10 g/L) and sonicated for 1 h to induce the partial breaking of BC to nanosized particles. Then, the so obtained BC suspension was centrifuged, and the supernatant was filtered (0.2 µm pore size membrane) and collected. A schematic representation of the BC NPs fabrication route is depicted in Figure 1.
As shown in Figure 2, BC NPs having an average size of ~2 nm are present in the filtered aqueous solution. This suspension was drop-casted onto the TiO2 support (Figure 1) with the aim of forming composite BC NPs-TiO2 materials.
Cross-section and top-view SEM images of TiO2 nanotube arrays used in this work are shown in Figure 3. The morphology of tubes appears well-ordered: an inner diameter of 104 ± 10 nm and a length of 5.6 ± 0.5 µm were measured over the entire surface of the photocatalyst. Such length, easily tunable by adjusting the anodization time, was selected as it proved to be a good compromise between high light harvesting and limited charge carrier recombination [49].
These one-dimensional TiO2 nanostructures were decorated with MBC or NBC nanoparticles produced by multiple sonication–centrifugation steps, as described in the Materials and Methods section. As shown in Figure 4a,e, BC nanoparticles (diameter < 25 nm) are present at the surface of the TiO2 nanotubes, already after a single drop-casting deposition (samples 1-MBC-TiO2 and 1-NBC-TiO2). This evidence suggests the suitability of the employed strategy for the production and deposition of BC nanoparticles.
As can be observed in Figure 4, the number and density of carbonaceous nanoparticles on the surface of the NTs arrays increase along with the number of depositions. This is particularly clear for the NBC-TiO2 samples (see Figure 4e–h), while for the MBC-TiO2 ones (Figure 4a–d) the effect is less visible; NBC-decorated NTs appear to be more heavily loaded with co-catalyst when compared to MBC-decorated ones. This evidence would indicate that both the number of drop-casting depositions and the nature of the feedstock heavily impact the BC loading on the photocatalysts.
No significant difference of the nanoparticle morphology was observed as a function of the number of depositions (see size distributions in Figure 5), since the shape of the BC co-catalysts remained nearly spherical. Only the nature of the feedstock seems to influence the BC NPs’ morphology; as depicted in Figure 5, NBC NPs are larger and exhibit a wider dimensional distribution (24.2 ± 7.8 nm) with respect to MBC ones (15.3 ± 4.0 nm). Such NP size indicates the occurrence of an agglomeration process during drop-casting deposition, as NPs suspended in the aqueous solution have a much lower dimension (~2 nm, see Figure 2). The difference between NBC and MBC NPs in terms of agglomeration degree can be ascribed to the diverse hydrophilicity of the materials. In detail, NBC is more hydrophobic than MBC, most likely due to the absence of a polar surface functional group [46]; according to the literature, the more hydrophobic BC NPs are, the higher their tendency to agglomerate in clusters is [50]. Such a consideration is also well in line with the much higher BC loading observed in NBC-TiO2 materials. As a confirmation of the occurrence of the agglomeration process, only a few distinct NPs are discernable when carrying out the seven drop-casting depositions (7-MBC-TiO2 and 7-NBC-TiO2); most of the deposited carbonaceous material clustered in a homogeneous thin film covering the surface of the TiO2 nanotubes (see Figure 4d,h).
Besides the investigation of compositional and morphological features of BC-TiO2 materials, XRD analysis (reported in Figure 6 and Figure S3) was performed to determine if the thermal treatment resulted in the crystallization of the amorphous TiO2 and which phases were formed. Diffraction peaks at 25.3°, 48°, 52.7°, and 59.2° were observed for the annealed sample: these peaks are attributed to the (101), (200), (105), and (211) planes of anatase TiO2. Additional peaks were observed for both the treated (annealed) and untreated samples at 38.3°, 38.8°, 40°, 52.8°, 70.5°, 76.6°, and 78°. These peaks can be attributed to the (002), (100), (101), (102), (110), (103), and (112) planes of hexagonal titanium, i.e., they can be associated with the metallic Ti support. Rutile formation was not observed, as, according to the literature, it would take place through nucleation at the oxide/metal substrate interface upon prolonged thermal treatments or at higher temperatures [51,52]. As expected, XRD performed on BC-decorated samples (Figure 6, green and red lines) showed that no TiO2 phase modifications are induced by the deposition of biochar. Moreover, no peaks that could be ascribed to the presence of crystalline carbon phases are visible, likely owing to the low amount and amorphous nature of the synthesized biochar.
Diffusive reflectance UV-vis spectra were also recorded for all the samples (Figure 7 and Figure S4) to investigate the potential effect of BC NPs in inducing a redshift in the light absorption spectrum from UV to the visible region: this feature is sometimes reported for bulk BC when used as supporting material for powdered TiO2 photocatalysts [36]. As expected, all the samples show high absorbance values for wavelengths lower than 380 nm, in line with the anatase TiO2 composition of the NTs. Moreover, periodic absorption signals are always present even in the undecorated TiO2 sample: this spectroscopic effect is most likely due to scattering inside the nanotube array [53] as already reported in previous works [54,55]. Interesting differences were found when decorating TiO2 with BC NPs. In detail, 5-NBC-TiO2 presents a broad and weak absorption peak in the 400–600 nm range, not present in any other sample, and the 5-MBC-TiO2 sample presents a peculiar peak with a relative maximum absorbance at 331 nm. The presence of this peak can be ascribed to the π-π* absorption typical of the aromatic system [56]; this assignment is also strengthened by the evidence of aromatic absorptions in the ATR spectrum recorded on such material [46]. Regarding the broad and weak absorption band present in the nutshell-derived sample, an analogous relatively weak absorbance in the visible region is widely reported in the literature [28,36,57,58], and this signal is generally attributed to the biochar content and to the interaction of the TiO2 with BC (Ti-O-C bonds). The low intensity of this absorption signal is reasonably ascribed to the low BC loading on the photocatalyst. Despite these differences in the UV absorption spectra, the energy band gaps (Eg), estimated using the Kubelka–Munk method (see inset in Figure 6) for TiO2 and BC-TiO2 materials, show no significant differences, ruling out the occurrence of shifts in the Eg, which remains in the UV light region even after decoration with BC NPs (Eg ~ 3.3 eV for all the samples).

2.3. Assessment of the Photocatalytic Performances

Prior to proceeding with photocatalytic experiments, MB stability under UV irradiation was preliminary carried out. A 10 mg/L dye solution was stirred and irradiated by UV light in the absence of any photocatalyst; as a result, no decrease of MB concentration was observed after three hours, demonstrating that MB does not undergo decomposition under these experimental conditions. Then, the presence of no light-assisted adsorption/degradation processes at the photocatalyst surface was investigated; tests were carried out by stirring the MB solution under dark conditions in the presence of bare TiO2 and BC-TiO2 photocatalysts. Even in these cases, no appreciable decreases of MB concentrations were observed for all tested catalysts after three hours using the bare or five times decorated materials (TiO2, 5-NBC-TiO2 and 5-MBC-TiO2). The absence of remarkable absorption phenomena can be ascribed to the low amount of BC deposited on the TiO2 nanotubes.
The photocatalytic tests were then performed to assess the effects of the BC NPs’ loading on the photocatalytic performances. It should be underlined that mild photocatalytic conditions were intentionally used to enhance differences between each system and to better analyze the kinetic of MB degradation; experiments under boosted conditions are discussed later.
A kinetic study over short-term experiments (three hours) was preliminary carried out. As shown in Figure 8, the presence of BC NPs significantly enhances the photocatalytic activity towards the degradation of MB, whatever their origin and amount. This evidence clearly demonstrates the suitability of using microalgae- and nutshell-derived BC NPs as co-catalysts.
Significant differences in photocatalytic efficiency were found by varying both the nature of the biomass feedstock and the loading of BC NPs on TiO2 nanotubes. For both systems, the degradation rate raises with the amount of co-catalyst (i.e., with the number of drop-casting depositions), reaching a maximum after five depositions (see Figure 8a,b); compared to undecorated TiO2, 2.5- and 2.8-times larger amounts of MB were removed after three hours by 5-MBC-TiO2 and 5-NBC-TiO2, respectively. When loading more carbonaceous material, i.e., by performing the drop-casting procedure seven times, a decrease in photocatalytic efficiency was observed (see samples 7-MBC-TiO2 and 7-NBC-TiO2 in Figure 8). These experimental findings are well in line with the typical bell-shaped behavior reported for the photocatalytic efficiency when varying the loading of co-catalysts [14].
Concerning the reaction pathway, since it was impossible to observe MB degradation via adsorption mechanism (i.e., in the absence of light irradiation) or photodecomposition of MB itself (i.e., in the absence of the photocatalyst), it appears clear that the MB degradation mechanism involves a photocatalytic process on the surface of BC-TiO2 materials. It is widely accepted that the overall process involves the formation of superoxide radicals, generated by a reaction with holes at the semiconductor surface, which quickly react with the dye molecules, breaking them down into smaller molecular fragments and eventually to CO2 and water [36]. This process is greatly enhanced by the sorption of MB on the BC NPs’ surfaces. Additionally, as metal particles (e.g., Pt) do when decorated on a semiconductor surface, biochar can shuttle and trap electrons [27]; this results in an improvement of charge carrier separation in the semiconductor, thus reducing the charge carrier recombination rate [28]. Based on the information present in literature, it can be reasonably supposed that the photocatalytic process occurring at the surface of the prepared materials is the combination of an adsorption process on BC nanoparticles and an oxidation reaction happening due to the presence of holes on the surface of the TiO2 nanotubes array. In this light, it is reasonable to suggest that the decomposition process is enhanced for samples which show isolated BC nanoparticles, while the formation of a BC film might lead to a decrease in photocatalytic activity due to a reduction of TiO2 surface available for the oxidation process. This interpretation explains the lower photocatalytic activity of 7-BC-TiO2 samples compared to that of 5-BC-TiO2 ones; in the former, a dense carbonaceous film completely covers the surface of TiO2 (where the oxidation of MB most likely takes place), whereas the latter shows well-defined BC nanoparticles and a more available TiO2 surface.
Based on these results, the most active materials, i.e., 5-NBC-TiO2 and 5-MBC-TiO2, were studied in view of their photocatalytic activity during longer experiments (up to 51 h) to assess their (photo) stability and gain insights into the degradation kinetics. MB degradation profiles are shown in Figure 9.
As can be observed for the long kinetics experiments, the MBC-decorated TiO2 is the most efficient photocatalyst for the degradation of MB, doubling the reaction rate compared to bare TiO2 (see Figure 10) and completely mineralizing the dye within 48 h. In contrast, undecorated and nutshell-derived materials were unable to completely degrade the azo dye and exhibited a less efficient behavior, even though 5-NBC-TiO2 performs even better than 5-MBC-TiO2 within the first 3 h (see Figure 8 and Figure 10). This behavior might be ascribed to a poisoning effect induced by methylene blue which is more marked for TiO2 and 5-NBC-TiO2 rather than for 5-MBC-TiO2. A second 3-h MB degradation was then performed on the catalysts used to assess if a partial deactivation could be observed. As can be seen in Figure 10, both TiO2 and 5-NBC-TiO2 show a marked deactivation, while 5-MBC-TiO2 shows only a slight decrease in its photocatalytic efficiency.
The reason behind the different behavior of the biochar materials can be ascribed to their different chemical composition; as reported in Section 2.1, the microalgae derived biochar presents a higher content in the oxygenated functional group and has a higher heteroatoms content altogether that results in a higher photocatalytic performance (likely due to a reduced recombination rate for electron–holes pairs). For the lignocellulosic-derived nanoparticles, it can instead be supposed that the enhanced photocatalyst activity is due to a higher MB adsorption compared to bare TiO2; however, the lack of oxygenated functional groups and heteroatoms results in an overall less active material.
Concerning the kinetic model, data fit with a first order model (Figure 9b); thus, we reasonably assumed that the mechanism is comparable to what is reported in the literature: MB molecules are adsorbed on BC, and MB degradation proceeds via reaction with superoxide radicals produced from the reaction between water and photogenerated holes [36].

2.4. Comparison of MB Degradation Performances with Literature Data

The two best performing materials (5-MBC-TiO2 and 5-NBC-TiO2) and pristine TiO2 NTs were tested under conditions closer to those most commonly reported in the literature for MB degradation, i.e., higher light power density and lower MB/catalyst ratios. These experiments were carried out to compare the performance of our materials against similar ones reported in the literature. It can be observed in Figure 11 that both MBC- and NBC-derived materials could completely degrade methylene blue within 3 h and did not exhibit a remarkable difference in photocatalytic performances.
A comparison of the performances of our materials against the BC-based ones reported in the literature is shown in Table 1. It should be underlined that, at present, only few articles report on MB degradation over BC-based photocatalysts. As a matter of fact, we did not find any data on BC-based planar photocatalysts, and thus only materials in powder form were considered for this comparison. In this respect, our systems represent the first examples of BC-decorated photocatalytic layers for this kind of application. Both microalgae- and nutshell-derived materials proposed in this work provide MB abatement rates comparable to or even higher than those reported for other BC-based photocatalysts. This is particularly interesting considering the much lower amount of co-catalyst present on our planar photocatalyst (in the range of micrograms) compared to that on reported materials. In any case, such an evaluation should be carefully considered owing to the quite different experimental conditions and photocatalyst form (i.e., planar platform vs. powder).
With the aim of further highlighting the competitive performances of BC-TiO2, different TiO2-based systems used under similar testing conditions were also considered in the comparison of the photocatalytic performances. Even when compared with these less green systems, the photocatalytic activity exhibited by TiO2 nanotube arrays decorated with BC NPs proposed in this work is comparable to or even higher than that reported for other TiO2-based photocatalysts.

3. Materials and Methods

3.1. TiO2 Nanotube Arrays Fabrication

Ti foils (99.945% purity, 0.03 mm thickness, William Gregor Ltd., London, UK) were degreased by sonicating in solvents of increasing polarity: acetone, isopropanol, ethanol, and ultrapure water. Each sonication step was conducted for 15 min and the samples were then dried under a N2 stream. Subsequently, cleaned Ti foils (15 × 15 mm) were anodized at room temperature in an ethylene glycol solution containing ultrapure water (2% v/v, produced by a Sartorius Arium mini UV Lab Water System) and a fluoride concentration of 0.09 M (obtained by dissolving NH4F, Merck, Kenilworth, NJ, USA). A two-electrode configuration was used for the anodization, where the Ti foil was the working electrode, and a graphite plate was the counter electrode. The anodization was performed by applying a potential of 40 V (for 60 min) using a DC power supply (Wanptek NPS1230W). The anodized samples were rinsed and soaked in ethanol for at least 30 min. The as-formed (amorphous) TiO2 nanotubes were then annealed and crystallized at 450 °C for 1 h in air with a heating ramp of 5 °C/min.

3.2. Biochar Production

Biochar was produced by slow pyrolysis at low temperature (350 °C) of two different commercially available feedstocks: nutshells and the microalgae Nannochloropsis sp. Biomasses were treated as detailed in previous work [46]. Briefly, approximately 10 g of dried raw biomass feedstock were put in a quartz tubular furnace and heated at 350 °C with a ramp of 10 °C/min under N2 stream (0.5 L/min). Isothermal conditions at 350 °C were kept for 1 h. The obtained biochar was pulverized using a Fritsch Pulverisette 0 agate ball mill. BC produced from nutshells and microalgae is labelled as NBC and MBC, respectively.

3.3. Fabrication of the Composite BC-TiO2 Photocatalyst

The fabrication of the composite BC-TiO2 photocatalysts is summarized in Figure 1.
After pyrolysis, the bulk BC powder was dispersed in ultrapure water (10 g/L) and sonicated for 1 h at room temperature using a Branson 5800 ultrasonic bath; this process induces the partial breaking of BC to nanosized particles [47,63]. The so obtained BC suspension was then centrifugated at 3000 rpm for 30 min. The supernatant was collected and centrifuged again to further disperse the smaller particles. Finally, the suspension obtained by the second centrifugation was filtered using a 0.2 µm pore size membrane; the collected liquid, containing nanosized BC, was deposited by drop-casting (120 µL) onto the TiO2 support. Then, the sample was dried under a laminar flow hood at room temperature for at least 6 h. The loading of BC was varied by performing multiple depositions (1, 3, 5, and 7 times): photocatalysts are accordingly labelled as n-MBC-TiO2 and n-NBC-TiO2, where “n” is the number of performed depositions.

3.4. Characterization of BC and BC-TiO2

The complete characterization of the morphology and chemical composition of BC (as obtained by pyrolysis) is reported in detail in our previous work [46].
The morphology of TiO2 NTs and BC-TiO2 was studied using an XL30 Environmental Scanning Electron Microscope (ESEM FEG Philips) at 20kV in high vacuum condition. Transmission electron microscopy (TEM) and high-resolution TEM (HR-TEM) analysis were performed by a ZEISS LIBRA 200FE microscope. A small volume (20 µL) of the suspension containing BC NPs was deposited onto a TEM grid to observe the morphology of as-formed NPs, i.e., before the deposition on the TiO2 nanotubes substrate. The particle-size distribution of BC nanoparticles was determined by processing SEM images using the software ImageJ (National Institutes of Health (NIH), Bethesda, MD, USA). X-ray diffraction (XRD) data were recorded on a Siemens D5000 system using Cu-Kα radiation, 40 kV, 40 mA, with a step of 0.03° (2θ) and a scanning speed of 0.06°·s−1. Diffuse reflectance UV-visible spectroscopy (DR-UV) measurement on solid samples and UV-vis analysis to determine the methylene blue concentration (in photocatalytic tests) were carried out using a Shimadzu UV-2600i spectrometer.

3.5. Photocatalytic Degradation of Methylene Blue

Methylene blue (MB) was selected as target molecule as this dye is widely studied and adopted as a pollutant model in the literature for the development of photocatalysts [27,43,64].
Photocatalytic tests under mild conditions were performed using a 250 mL quartz beaker placed on a magnetic stirrer to ensure the continuous mixing of the solution. Each experiment was performed using 100 mL of a 10 mg/L MB solution (obtained by dilution of a concentrated 1 g/L solution prepared by dissolving solid MB in ultrapure water) and an irradiation power of 40 mW/cm2.
Boosted photocatalytic tests were performed using 10 mL of a 10 mg/L MB solution stirred in a quartz test tube with an irradiation power of 100 mW/cm2.
An UV LED light source (λ = 365 nm, UVWave) with tunable power was used for both types of photocatalytic tests. Power density was measured by means of a thermal sensor (OptoSigma®).
MB concentration was monitored during the degradation tests by spectrophotometric means, recording absorbance values at 665 nm. A 30-min equilibration without light irradiation was performed before each photocatalytic test.

4. Conclusions

In this paper we demonstrated the feasibility of producing biochar-decorated TiO2 photocatalysts by means of simple drop-casting. The composite BC-TiO2 materials prepared in the present work proved to be more efficient than undecorated TiO2 nanotubes towards the degradation of methylene blue under UV light irradiation; this evidence clearly indicates the ability of BC NPs to act as co-catalysts on the TiO2 surface. The greatly enhanced sorption of organic molecules, together with improved charge carrier separation provided by BC NPs led to almost tripled photocatalytic performances.
It was observed that two types of biochar derived from markedly different biomasses (i.e., microalgae and nutshells) can promote MB decomposition to a similar extent, even if microalgae-derived BC is much less prone to deactivation than the nutshell-derived one. Further studies will be carried out to deeply understand the nature of the deactivation process and restore the original activity of the material.
Future perspectives might involve the optimization of reaction conditions to obtain the best photocatalytic performances and the application of biochar nanoparticle-decorated photocatalysts for the abatement of contaminants of emerging concern (e.g., pharmaceuticals such as acetaminophen).

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/catal11091048/s1, Figure S1: SEM images of fresh biomass and bulk biochar, Figure S2: elemental composition of biochar, Figure S3: XRD patterns of MBC- and NBC-TiO2, Figure S4: DR-UV spectra of MBC- and NBC-TiO2.

Author Contributions

Conceptualization, D.S., D.M., and S.R.; methodology, M.P. and D.S.; investigation, M.P., D.S., G.B., and M.M.; formal analysis, M.P.; validation, M.P.; data curation, M.P., D.S., and G.B.; writing—original draft preparation, M.P. and D.S.; writing—review and editing, D.M., S.R., C.D., and M.A.; visualization, D.S. and G.B.; resources, C.D. and S.R.; supervision, C.D., D.M., and S.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Liyanage, C.P.; Yamada, K. Impact of Population Growth on the Water Quality of Natural Water Bodies. Sustainability 2017, 9, 1405. [Google Scholar] [CrossRef] [Green Version]
  2. Mills, A.; Davies, R.H.; Worsley, D. Water purification by semiconductor photocatalysis. Chem. Soc. Rev. 1993, 22, 417. [Google Scholar] [CrossRef]
  3. Gopinath, K.P.; Madhav, N.V.; Krishnan, A.; Malolan, R.; Rangarajan, G. Present applications of titanium dioxide for the photocatalytic removal of pollutants from water: A review. J. Environ. Manag. 2020, 270, 110906. [Google Scholar] [CrossRef]
  4. Zhao, Z.; An, H.; Lin, J.; Feng, M.; Murugadoss, V.; Ding, T.; Liu, H.; Shao, Q.; Mai, X.; Wang, N.; et al. Progress on the Photocatalytic Reduction Removal of Chromium Contamination. Chem. Rec. 2019, 19, 873–882. [Google Scholar] [CrossRef]
  5. Spanu, D.; Dal Santo, V.; Malara, F.; Naldoni, A.; Turolla, A.; Antonelli, M.; Dossi, C.; Marelli, M.; Altomare, M.; Schmuki, P.; et al. Photoelectrocatalytic oxidation of As(III) over hematite photoanodes: A sensible indicator of the presence of highly reactive surface sites. Electrochim. Acta 2018, 292, 828–837. [Google Scholar] [CrossRef]
  6. Spanu, D.; Bestetti, A.; Hildebrand, H.; Schmuki, P.; Altomare, M.; Recchia, S. Photocatalytic reduction and scavenging of Hg(II) over templated-dewetted Au on TiO2 nanotubes. Photochem. Photobiol. Sci. 2019, 18, 1046–1055. [Google Scholar] [CrossRef]
  7. Chatterjee, D.; Dasgupta, S. Visible light induced photocatalytic degradation of organic pollutants. J. Photochem. Photobiol. C Photochem. Rev. 2005, 6, 186–205. [Google Scholar] [CrossRef]
  8. Gusain, R.; Gupta, K.; Joshi, P.; Khatri, O.P. Adsorptive removal and photocatalytic degradation of organic pollutants using metal oxides and their composites: A comprehensive review. Adv. Colloid Interface Sci. 2019, 272, 102009. [Google Scholar] [CrossRef] [PubMed]
  9. Hashimoto, K.; Irie, H.; Fujishima, A. TiO2 Photocatalysis: A Historical Overview and Future Prospects. Jpn. J. Appl. Phys. 2005, 44, 8269–8285. [Google Scholar] [CrossRef]
  10. Nakata, K.; Fujishima, A. TiO2 photocatalysis: Design and applications. J. Photochem. Photobiol. C Photochem. Rev. 2012, 13, 169–189. [Google Scholar] [CrossRef]
  11. Al-Azri, Z.H.N.; Chen, W.-T.; Chan, A.; Jovic, V.; Ina, T.; Idriss, H.; Waterhouse, G.I.N. The roles of metal co-catalysts and reaction media in photocatalytic hydrogen production: Performance evaluation of M/TiO2 photocatalysts (M = Pd, Pt, Au) in different alcohol–water mixtures. J. Catal. 2015, 329, 355–367. [Google Scholar] [CrossRef]
  12. Spanu, D.; Recchia, S.; Mohajernia, S.; Schmuki, P.; Altomare, M. Site-selective Pt dewetting on WO3-coated TiO2 nanotube arrays: An electron transfer cascade-based H2 evolution photocatalyst. Appl. Catal. B Environ. 2018, 237, 198–205. [Google Scholar] [CrossRef]
  13. Meng, A.; Zhang, L.; Cheng, B.; Yu, J. Dual Cocatalysts in TiO2 Photocatalysis. Adv. Mater. 2019, 31, 1807660. [Google Scholar] [CrossRef]
  14. Spanu, D.; Minguzzi, A.; Recchia, S.; Shahvardanfard, F.; Tomanec, O.; Zboril, R.; Schmuki, P.; Ghigna, P.; Altomare, M. An Operando X-ray Absorption Spectroscopy Study of a NiCu−TiO2 Photocatalyst for H2 Evolution. ACS Catal. 2020, 10, 8293–8302. [Google Scholar] [CrossRef]
  15. Iqbal, A.; Kafizas, A.; Sotelo-Vazquez, C.; Wilson, R.; Ling, M.; Taylor, A.; Blackman, C.; Bevan, K.; Parkin, I.; Quesada-Cabrera, R. Charge Transport Phenomena in Heterojunction Photocatalysts: The WO3/TiO2 System as an Archetypical Model. ACS Appl. Mater. Interfaces 2021, 13, 9781–9793. [Google Scholar] [CrossRef] [PubMed]
  16. Mohamed, H.H.; Alomair, N.A.; Akhtar, S.; Youssef, T.E. Eco-friendly synthesized α-Fe2O3/TiO2 heterojunction with enhanced visible light photocatalytic activity. J. Photochem. Photobiol. A Chem. 2019, 382, 111951. [Google Scholar] [CrossRef]
  17. Altomare, M.; Nguyen, N.T.; Hejazi, S.; Schmuki, P. A Cocatalytic Electron-Transfer Cascade Site-Selectively Placed on TiO2 Nanotubes Yields Enhanced Photocatalytic H2 Evolution. Adv. Funct. Mater. 2018, 28, 1704259. [Google Scholar] [CrossRef]
  18. Ji, L.; Spanu, D.; Denisov, N.; Recchia, S.; Schmuki, P.; Altomare, M. A Dewetted-Dealloyed Nanoporous Pt Co-Catalyst Formed on TiO2 Nanotube Arrays Leads to Strongly Enhanced Photocatalytic H2 Production. Chem. An Asian J. 2020, 15, 301–309. [Google Scholar] [CrossRef] [Green Version]
  19. Spanu, D.; Recchia, S.; Mohajernia, S.; Tomanec, O.; Kment, Š.; Zboril, R.; Schmuki, P.; Altomare, M. Templated Dewetting–Alloying of NiCu Bilayers on TiO2 Nanotubes Enables Efficient Noble-Metal-Free Photocatalytic H2 Evolution. ACS Catal. 2018, 8, 5298–5305. [Google Scholar] [CrossRef] [Green Version]
  20. Zielińska-Jurek, A. Progress, Challenge, and Perspective of Bimetallic TiO2 -Based Photocatalysts. J. Nanomater. 2014, 2014, 208920. [Google Scholar] [CrossRef] [Green Version]
  21. Woan, K.; Pyrgiotakis, G.; Sigmund, W. Photocatalytic Carbon-Nanotube-TiO2 Composites. Adv. Mater. 2009, 21, 2233–2239. [Google Scholar] [CrossRef]
  22. Yu, Y.; Yu, J.C.; Yu, J.-G.; Kwok, Y.-C.; Che, Y.-K.; Zhao, J.-C.; Ding, L.; Ge, W.-K.; Wong, P.-K. Enhancement of photocatalytic activity of mesoporous TiO2 by using carbon nanotubes. Appl. Catal. A Gen. 2005, 289, 186–196. [Google Scholar] [CrossRef]
  23. Minella, M.; Sordello, F.; Minero, C. Photocatalytic process in TiO2/graphene hybrid materials. Evidence of charge separation by electron transfer from reduced graphene oxide to TiO2. Catal. Today 2017, 281, 29–37. [Google Scholar] [CrossRef]
  24. Chen, P.; Wang, F.; Chen, Z.-F.; Zhang, Q.; Su, Y.; Shen, L.; Yao, K.; Liu, Y.; Cai, Z.; Lv, W.; et al. Study on the photocatalytic mechanism and detoxicity of gemfibrozil by a sunlight-driven TiO2/carbon dots photocatalyst: The significant roles of reactive oxygen species. Appl. Catal. B Environ. 2017, 204, 250–259. [Google Scholar] [CrossRef]
  25. Wang, Q.; Huang, J.; Sun, H.; Zhang, K.Q.; Lai, Y. Uniform carbon dots@TiO2 nanotube arrays with full spectrum wavelength light activation for efficient dye degradation and overall water splitting. Nanoscale 2017, 9, 16046–16058. [Google Scholar] [CrossRef] [PubMed]
  26. Colmenares, J.C.; Varma, R.S.; Lisowski, P. Sustainable hybrid photocatalysts: Titania immobilized on carbon materials derived from renewable and biodegradable resources. Green Chem. 2016, 18, 5736–5750. [Google Scholar] [CrossRef] [Green Version]
  27. Mian, M.M.; Liu, G. Recent progress in biochar-supported photocatalysts: Synthesis, role of biochar, and applications. RSC Adv. 2018, 8, 14237–14248. [Google Scholar] [CrossRef] [Green Version]
  28. Cui, J.; Zhang, F.; Li, H.; Cui, J.; Ren, Y.; Yu, X. Recent progress in biochar-based photocatalysts for wastewater treatment: Synthesis, mechanisms, and applications. Appl. Sci. 2020, 10, 1019. [Google Scholar] [CrossRef] [Green Version]
  29. Chen, Q.; Lin, G.; Meng, L.; Zhou, L.; Hu, L.; Nong, J.; Li, Y.; Wang, J.; Hu, K.; Yu, Q. Enhanced photoelectric performance of TiO2 nanotubes sensitized with carbon dots derived from bagasse. Chem. Phys. Lett. 2020, 749, 137428. [Google Scholar] [CrossRef]
  30. Guo, D.; Li, Y.; Cui, B.; Hu, M.; Luo, S.; Ji, B.; Liu, Y. Natural adsorption of methylene blue by waste fallen leaves of Magnoliaceae and its repeated thermal regeneration for reuse. J. Clean. Prod. 2020, 267, 121903. [Google Scholar] [CrossRef]
  31. Ji, B.; Zhu, L.; Wang, S.; Liu, Y. Temperature-effect on the performance of non-aerated microalgal-bacterial granular sludge process in municipal wastewater treatment. J. Environ. Manag. 2021, 282, 111955. [Google Scholar] [CrossRef]
  32. González-Castaño, M.; Morales, C.; Navarro de Miguel, J.C.; Boelte, J.H.; Klepel, O.; Flege, J.I.; Arellano-García, H. Are Ni/and Ni5Fe1/biochar catalysts suitable for synthetic natural gas production? A comparison with γ-Al2O3 supported catalysts. Green Energy Environ. 2021. [Google Scholar] [CrossRef]
  33. Kant Bhatia, S.; Palai, A.K.; Kumar, A.; Kant Bhatia, R.; Kumar Patel, A.; Kumar Thakur, V.; Yang, Y.-H. Trends in renewable energy production employing biomass-based biochar. Bioresour. Technol. 2021, 340, 125644. [Google Scholar] [CrossRef]
  34. Spanu, D.; Binda, G.; Dossi, C.; Monticelli, D. Biochar as an alternative sustainable platform for sensing applications: A review. Microchem. J. 2020, 159, 105506. [Google Scholar] [CrossRef]
  35. Gibertini, E.; Liberale, F.; Dossi, C.; Binda, G.; Mattioli, B.; Bettinetti, R.; Maspero, A.; Fiore, M.; Ruffo, R.; Magagnin, L. Algae-derived hard carbon anodes for Na-ion batteries. J. Appl. Electrochem. 2021. [Google Scholar] [CrossRef]
  36. Fazal, T.; Razzaq, A.; Javed, F.; Hafeez, A.; Rashid, N.; Amjad, U.S.; Ur Rehman, M.S.; Faisal, A.; Rehman, F. Integrating adsorption and photocatalysis: A cost effective strategy for textile wastewater treatment using hybrid biochar-TiO2 composite. J. Hazard. Mater. 2020, 390, 121623. [Google Scholar] [CrossRef]
  37. Zhao, L.; Cao, X.; Mašek, O.; Zimmerman, A. Heterogeneity of biochar properties as a function of feedstock sources and production temperatures. J. Hazard. Mater. 2013, 256–257, 1–9. [Google Scholar] [CrossRef] [Green Version]
  38. Kloss, S.; Zehetner, F.; Dellantonio, A.; Hamid, R.; Ottner, F.; Liedtke, V.; Schwanninger, M.; Gerzabek, M.H.; Soja, G. Characterization of Slow Pyrolysis Biochars: Effects of Feedstocks and Pyrolysis Temperature on Biochar Properties. J. Environ. Qual. 2012, 41, 990–1000. [Google Scholar] [CrossRef] [PubMed]
  39. Tomczyk, A.; Sokołowska, Z.; Boguta, P. Biochar physicochemical properties: Pyrolysis temperature and feedstock kind effects. Rev. Environ. Sci. Bio/Technol. 2020, 19, 191–215. [Google Scholar] [CrossRef] [Green Version]
  40. Wang, W.; Zhang, J.; Chen, T.; Sun, J.; Ma, X.; Wang, Y.; Wang, J.; Xie, Z. Preparation of TiO2-modified Biochar and its Characteristics of Photo-catalysis Degradation for Enrofloxacin. Sci. Rep. 2020, 10, 6588. [Google Scholar] [CrossRef]
  41. Angın, D.; Altintig, E.; Köse, T.E. Influence of process parameters on the surface and chemical properties of activated carbon obtained from biochar by chemical activation. Bioresour. Technol. 2013, 148, 542–549. [Google Scholar] [CrossRef]
  42. Ren, Q.; Wu, Z.; Hu, S.; He, L.; Su, S.; Wang, Y.; Jiang, L.; Xiang, J. Sulfur self-doped char with high specific capacitance derived from waste tire: Effects of pyrolysis temperature. Sci. Total Environ. 2020, 741, 140193. [Google Scholar] [CrossRef]
  43. Zhu, Y.; Yi, B.; Yuan, Q.; Wu, Y.; Wang, M.; Yan, S. Removal of methylene blue from aqueous solution by cattle manure-derived low temperature biochar. RSC Adv. 2018, 8, 19917–19929. [Google Scholar] [CrossRef] [Green Version]
  44. Xu, X.; Cao, X.; Zhao, L.; Wang, H.; Yu, H.; Gao, B. Removal of Cu, Zn, and Cd from aqueous solutions by the dairy manure-derived biochar. Environ. Sci. Pollut. Res. 2013, 20, 358–368. [Google Scholar] [CrossRef]
  45. Fernandes, J.O.; Bernardino, C.A.R.; Mahler, C.F.; Santelli, R.E.; Braz, B.F.; Borges, R.C.; da Cunha Veloso, M.C.; Romeiro, G.A.; Cincotto, F.H. Biochar Generated from Agro-Industry Sugarcane Residue by Low Temperature Pyrolysis Utilized as an Adsorption Agent for the Removal of Thiamethoxam Pesticide in Wastewater. Water Air Soil Pollut. 2021, 232, 67. [Google Scholar] [CrossRef]
  46. Binda, G.; Spanu, D.; Bettinetti, R.; Magagnin, L.; Pozzi, A.; Dossi, C. Comprehensive comparison of microalgae-derived biochar from different feedstocks: A prospective study for future environmental applications. Algal Res. 2020, 52, 102103. [Google Scholar] [CrossRef]
  47. Plácido, J.; Bustamante-López, S.; Meissner, K.E.; Kelly, D.E.; Kelly, S.L. Microalgae biochar-derived carbon dots and their application in heavy metal sensing in aqueous systems. Sci. Total Environ. 2019, 656, 531–539. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Liu, G.; Zheng, H.; Jiang, Z.; Zhao, J.; Wang, Z.; Pan, B.; Xing, B. Formation and Physicochemical Characteristics of Nano Biochar: Insight into Chemical and Colloidal Stability. Environ. Sci. Technol. 2018, 52, 10369–10379. [Google Scholar] [CrossRef]
  49. Cha, G.; Schmuki, P.; Altomare, M. Anodic TiO2 nanotube membranes: Site-selective Pt-activation and photocatalytic H2 evolution. Electrochim. Acta 2017, 258, 302–310. [Google Scholar] [CrossRef]
  50. Tansel, B.; Boglaienko, D. Characterization of aggregation and declustering tendency of hydrophobic fine particles in water. Granul. Matter. 2019, 21, 34. [Google Scholar] [CrossRef]
  51. Albu, S.P.; Tsuchiya, H.; Fujimoto, S.; Schmuki, P. TiO2 Nanotubes—Annealing Effects on Detailed Morphology and Structure. Eur. J. Inorg. Chem. 2010, 2010, 4351–4356. [Google Scholar] [CrossRef]
  52. Altomare, M.; Pozzi, M.; Allieta, M.; Bettini, L.G.; Selli, E. H2 and O2 photocatalytic production on TiO2 nanotube arrays: Effect of the anodization time on structural features and photoactivity. Appl. Catal. B Environ. 2013, 136–137, 81–88. [Google Scholar] [CrossRef]
  53. Abdelmoula, M.; Sokoloff, J.; Lu, W.-T.; Close, T.; Menon, L.; Richter, C. Optical properties of titanium dioxide nanotube arrays. J. Appl. Phys. 2014, 115, 014306. [Google Scholar] [CrossRef]
  54. Olgun, U.; Gülfen, M.; Üstel, F.; Arslan, H. Electro-Optics and Band Gap Energies of Nanosilver-Coated TiO2 Nanotubes on Titanium Metal. Acta Metall. Sin. Engl. Lett. 2018, 31, 153–163. [Google Scholar] [CrossRef] [Green Version]
  55. Li, J.; Wang, Y.; Wang, L. Structure and properties of nitrogen incorporated in TiO2 nanotubes array. Mater. Res. Express 2014, 1, 025040. [Google Scholar] [CrossRef]
  56. Helwig, A.; Müller, G.; Paul, S. Health monitoring of aviation hydraulic fluids using opto-chemical sensor technologies. Chemosensors 2020, 8, 131. [Google Scholar] [CrossRef]
  57. Pang, Y.L.; Law, Z.X.; Lim, S.; Chan, Y.Y.; Shuit, S.H.; Chong, W.C.; Lai, C.W. Enhanced photocatalytic degradation of methyl orange by coconut shell–derived biochar composites under visible LED light irradiation. Environ. Sci. Pollut. Res. 2021, 28, 27457–27473. [Google Scholar] [CrossRef]
  58. Xie, X.; Li, S.; Zhang, H.; Wang, Z.; Huang, H. Promoting charge separation of biochar-based Zn-TiO2/pBC in the presence of ZnO for efficient sulfamethoxazole photodegradation under visible light irradiation. Sci. Total Environ. 2019, 659, 529–539. [Google Scholar] [CrossRef]
  59. Silvestri, S.; Stefanello, N.; Sulkovski, A.A.; Foletto, E.L. Preparation of TiO2 supported on MDF biochar for simultaneous removal of methylene blue by adsorption and photocatalysis. J. Chem. Technol. Biotechnol. 2020, 95, 2723–2729. [Google Scholar] [CrossRef]
  60. Matos, J. Eco-Friendly Heterogeneous Photocatalysis on Biochar-Based Materials under Solar Irradiation. Top. Catal. 2016, 59, 394–402. [Google Scholar] [CrossRef]
  61. Ciobanu, V.; Plesco, I. TiO2 nanotubes for photocatalytic degradation of methylene blue. J. Eng. Sci. 2021, 28, 23–30. [Google Scholar] [CrossRef]
  62. Song, Y.S.; Kim, Y.; Lee, H.; Lee, D.Y.; Lee, M.H.; Kim, B.Y. Photocatalytic Activity of Al-TiO2 Nanomaterials for the Degradation of Methylene Blue Dye. J. Korean Phys. Soc. 2018, 72, 412–416. [Google Scholar] [CrossRef]
  63. Meng, W.; Bai, X.; Wang, B.; Liu, Z.; Lu, S.; Yang, B. Biomass-Derived Carbon Dots and Their Applications. Energy Environ. Mater. 2019, 2, 172–192. [Google Scholar] [CrossRef]
  64. Abdellah, M.H.; Nosier, S.A.; El-Shazly, A.H.; Mubarak, A.A. Photocatalytic decolorization of methylene blue using TiO2/UV system enhanced by air sparging. Alexandria Eng. J. 2018, 57, 3727–3735. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of the steps involved in the fabrication of BC-TiO2 photocatalysts.
Figure 1. Schematic representation of the steps involved in the fabrication of BC-TiO2 photocatalysts.
Catalysts 11 01048 g001
Figure 2. TEM images of (a) NBC and (b) MBC NPs produced by performing the procedure depicted in Figure 1 and described in the Materials and Methods section.
Figure 2. TEM images of (a) NBC and (b) MBC NPs produced by performing the procedure depicted in Figure 1 and described in the Materials and Methods section.
Catalysts 11 01048 g002
Figure 3. SEM images of (a) cross-section and (b) top-view of bare TiO2 nanotubes (inset: magnified top-view image).
Figure 3. SEM images of (a) cross-section and (b) top-view of bare TiO2 nanotubes (inset: magnified top-view image).
Catalysts 11 01048 g003
Figure 4. SEM top-view images of the samples obtained by depositing microalgae- (ad) and nutshell-derived (eh) BC NPs on the surface of TiO2 nanotube arrays. Different BC loadings were achieved by performing multiple drop-casting depositions: samples after (a,e) 1, (b,f) 3, (c,g) 5, and (d,h) 7 depositions are depicted. Insets: magnified SEM top-view images.
Figure 4. SEM top-view images of the samples obtained by depositing microalgae- (ad) and nutshell-derived (eh) BC NPs on the surface of TiO2 nanotube arrays. Different BC loadings were achieved by performing multiple drop-casting depositions: samples after (a,e) 1, (b,f) 3, (c,g) 5, and (d,h) 7 depositions are depicted. Insets: magnified SEM top-view images.
Catalysts 11 01048 g004
Figure 5. Size distributions of BC NPs in samples after (a,d) 1, (b,e) 3, (c,f) and 5 depositions. 7-NBC-TiO2 and 7-MBC-TiO2 samples were not considered in this morphological evaluation owing to the impossibility of identifying well-defined NPs due to the formation of agglomerated carbonaceous film. Refer to Figure 4 for SEM images of these samples.
Figure 5. Size distributions of BC NPs in samples after (a,d) 1, (b,e) 3, (c,f) and 5 depositions. 7-NBC-TiO2 and 7-MBC-TiO2 samples were not considered in this morphological evaluation owing to the impossibility of identifying well-defined NPs due to the formation of agglomerated carbonaceous film. Refer to Figure 4 for SEM images of these samples.
Catalysts 11 01048 g005
Figure 6. XRD pattern of as-formed (blue line) and crystalline TiO2 nanotubes (black line), 5-MBC-TiO2 (green line), and 5-NBC-TiO2 (red line). The peak labels are indicated as follows: A = Anatase, Ti = Titanium (substrate), Al = Aluminum (sample holder).
Figure 6. XRD pattern of as-formed (blue line) and crystalline TiO2 nanotubes (black line), 5-MBC-TiO2 (green line), and 5-NBC-TiO2 (red line). The peak labels are indicated as follows: A = Anatase, Ti = Titanium (substrate), Al = Aluminum (sample holder).
Catalysts 11 01048 g006
Figure 7. DR-UV spectra of bare TiO2 nanotubes (black line), 5-NBC-TiO2 (red line), and 5-MBC-TiO2 (green line). Inset: Kubelka–Munk plots for the samples.
Figure 7. DR-UV spectra of bare TiO2 nanotubes (black line), 5-NBC-TiO2 (red line), and 5-MBC-TiO2 (green line). Inset: Kubelka–Munk plots for the samples.
Catalysts 11 01048 g007
Figure 8. Effect of BC loading on the degradation of MB by nutshell—(a) and microalgae—(b) derived photocatalysts.
Figure 8. Effect of BC loading on the degradation of MB by nutshell—(a) and microalgae—(b) derived photocatalysts.
Catalysts 11 01048 g008
Figure 9. (a) Fifty-one-hour degradation tests performed on TiO2 (black), 5-NBC-TiO2 (red), and 5-MBC-TiO2 (green). (b) Linearization by fitting with a first order kinetic model.
Figure 9. (a) Fifty-one-hour degradation tests performed on TiO2 (black), 5-NBC-TiO2 (red), and 5-MBC-TiO2 (green). (b) Linearization by fitting with a first order kinetic model.
Catalysts 11 01048 g009
Figure 10. Kinetic constants for fresh and recycled photocatalysts: all values are calculated from the data collected within the first 3 h of reaction under mild conditions.
Figure 10. Kinetic constants for fresh and recycled photocatalysts: all values are calculated from the data collected within the first 3 h of reaction under mild conditions.
Catalysts 11 01048 g010
Figure 11. Kinetic curve obtained for MB photocatalytic degradation using bare TiO2 (black line), 5-MBC-TiO2 (green line), and 5-NBC-TiO2 (red line) under boosted conditions.
Figure 11. Kinetic curve obtained for MB photocatalytic degradation using bare TiO2 (black line), 5-MBC-TiO2 (green line), and 5-NBC-TiO2 (red line) under boosted conditions.
Catalysts 11 01048 g011
Table 1. Comparison of the photocatalytic performances towards MB degradation achieved in this work with those reported in the literature for other TiO2-BC composite materials.
Table 1. Comparison of the photocatalytic performances towards MB degradation achieved in this work with those reported in the literature for other TiO2-BC composite materials.
PhotocatalystBiomass FeedstockExperimental
Conditions
% of MB Consumed in 2 hTime for Complete MB DegradationReference
TiO2 nanotube arrays decorated with BC NPsMicroalgae
(Nannochloropsis sp.)
(i) 10 mL of 10 mg/L MB;
(ii) 100 mW/cm2 UV light (365 nm)
90%3 hThis work
TiO2 nanotube arrays decorated with BC NPsNutshell(i) 10 mL of 10 mg/L MB;
(ii) 100 mW/cm2 UV light (365 nm)
90%3 hThis work
BC impregnated with TiO2 (10% wt.)Medium density fireboard(i) 0.4 g/L of photocatalyst in 100 mg/L MB;
(ii) 10 mW/m2 UV light (365 nm)
60%n.a.[59]
BC decorated with TiO2 NPsMacroalgae(i) 1 g/L of photocatalyst in 5 mg/L in MB solution
(ii) 500 W visible light bulb
90%3 h[36]
BC-TiO2 hybrid material (10% wt.)Plum stones(i) 0.56 g/L of photocatalyst in 25 mg/L MB;
(ii) Metal halide lamp (522.7 W/m2)
≤50%≥4 h[60]
Powdered TiO2 nanotubesn.a.(i) 0.4 g/L of photocatalyst in 31 mg/L MB;
(ii) 100 mW/cm2 UV light (365 nm)
85%n.a.[61]
Powdered Al-TiO2 nanotubesn.a.(i) 0.2 g/L of photocatalyst in 10 mg/L MB;
(ii) 100 mW/cm2 UV light (365 nm)
90%n.a.[62]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Pinna, M.; Binda, G.; Altomare, M.; Marelli, M.; Dossi, C.; Monticelli, D.; Spanu, D.; Recchia, S. Biochar Nanoparticles over TiO2 Nanotube Arrays: A Green Co-Catalyst to Boost the Photocatalytic Degradation of Organic Pollutants. Catalysts 2021, 11, 1048. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11091048

AMA Style

Pinna M, Binda G, Altomare M, Marelli M, Dossi C, Monticelli D, Spanu D, Recchia S. Biochar Nanoparticles over TiO2 Nanotube Arrays: A Green Co-Catalyst to Boost the Photocatalytic Degradation of Organic Pollutants. Catalysts. 2021; 11(9):1048. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11091048

Chicago/Turabian Style

Pinna, Marco, Gilberto Binda, Marco Altomare, Marcello Marelli, Carlo Dossi, Damiano Monticelli, Davide Spanu, and Sandro Recchia. 2021. "Biochar Nanoparticles over TiO2 Nanotube Arrays: A Green Co-Catalyst to Boost the Photocatalytic Degradation of Organic Pollutants" Catalysts 11, no. 9: 1048. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11091048

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