ACS Publications. Most Trusted. Most Cited. Most Read
My Activity
CONTENT TYPES

Figure 1Loading Img

Combined Structural and Plasmonic Enhancement of Nanometer-Thin Film Photocatalysis for Solar-Driven Wastewater Treatment

  • Desislava Daskalova
    Desislava Daskalova
    Advanced Microelectronic Center Aachen, AMO GmbH, 52074 Aachen, Germany
    Chair of Electronic Devices, RWTH Aachen University, 52074 Aachen, Germany
  • Gonzalo Aguila Flores
    Gonzalo Aguila Flores
    Advanced Microelectronic Center Aachen, AMO GmbH, 52074 Aachen, Germany
  • Ulrich Plachetka
    Ulrich Plachetka
    Advanced Microelectronic Center Aachen, AMO GmbH, 52074 Aachen, Germany
  • Michael Möller
    Michael Möller
    Advanced Microelectronic Center Aachen, AMO GmbH, 52074 Aachen, Germany
  • Julia Wolters
    Julia Wolters
    Institute of Environmental Engineering, RWTH Aachen University, 52074 Aachen, Germany
  • Thomas Wintgens
    Thomas Wintgens
    Institute of Environmental Engineering, RWTH Aachen University, 52074 Aachen, Germany
  • , and 
  • Max C. Lemme*
    Max C. Lemme
    Advanced Microelectronic Center Aachen, AMO GmbH, 52074 Aachen, Germany
    Chair of Electronic Devices, RWTH Aachen University, 52074 Aachen, Germany
    *Email: [email protected]. Phone: +49 2418867200.
    More by Max C. Lemme
Cite this: ACS Appl. Nano Mater. 2023, 6, 16, 15204–15212
Publication Date (Web):August 16, 2023
https://doi.org/10.1021/acsanm.3c02867

Copyright © 2022 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY-NC-ND 4.0.
  • Open Access

Article Views

427

Altmetric

-

Citations

-
LEARN ABOUT THESE METRICS
PDF (4 MB)
Supporting Info (1)»

Abstract

Titanium dioxide (TiO2) thin films are commonly used as photocatalytic materials. Here, we enhance the photocatalytic activity of devices based on titanium dioxide (TiO2) by combining nanostructured glass substrates with metallic plasmonic nanostructures. We achieve a three-fold increase of the catalyst’s surface area through nanoscale, three-dimensional patterning of periodic, conical grids, which creates a broadband optical absorber. The addition of aluminum and gold activates the structures plasmonically and increases the optical absorption in the TiO2 films to above 70% in the visible and NIR spectral range. We demonstrate the resulting enhancement of the photocatalytic activity with organic dye degradation tests under different light sources. Furthermore, the pharmaceutical drug Carbamazepine, a common water pollutant, is reduced in the aqueous solution by up to 48% in 360 min. Our approach is scalable and potentially enables future solar-driven wastewater treatment.

This publication is licensed under

CC-BY-NC-ND 4.0.
  • cc licence
  • by licence
  • nc licence
  • nd licence

Introduction

ARTICLE SECTIONS
Jump To

The use of titanium dioxide (TiO2) nanoparticles or porous films is well-established in heterogeneous photocatalysis. Under ultraviolet (UV) light irradiation, TiO2 absorbs highly energetic photons, generating electron–hole pairs. These photoexcited charge carriers can then migrate to the surface of the semiconductor to participate in advanced oxidation processes. Owing to their large surface-to-volume ratio, TiO2 nanoparticles in different modifications have been demonstrated to possess excellent photocatalytic performance at short wavelengths. (1) However, certain applications like wastewater treatment require stable TiO2 films instead of nanoparticles because of the inherent difficulties to filter and reuse the nanoparticles. (2) In addition, introducing large amounts of engineered nanomaterials into the biosphere poses toxicological and environmental risks. (3)
As an alternative, the advancement and wide availability of thin film technologies enable the application of thin film photocatalysis on an industrial scale. Nevertheless, thin film photocatalysts lack in efficiency compared to nanoparticles due to their lower surface area and fewer reactive sites. Hence, different strategies have been employed to improve their performance, such as thermally induced nanocrack networks in sputtered TiO2 films to enhance the surface area, (4) or semiconductor heterojunctions to enhance photocatalytic activity under UV (5,6) and extend the absorption to the visible range. (7) Utilizing plasmonic nanostructures to enhance photocatalysis is another promising route toward the commercialization of solar-driven photocatalysis. (8,9) It is based on the collective oscillations of electrons in the metal under illumination (plasmons), which generate intense and strongly localized electric fields in the vicinity of the metallic nanostructures. These plasmonic “hot spots”, when coupled to a semiconducting photocatalyst, can then aid photocatalysis via effects like strong light absorption and scattering, electromagnetic field enhancement, and hot carrier generation. While its fundamentals are well understood, the interplay between different mechanisms of enhancement, like direct electron transfer and plasmonic resonance electron transfer at the metal–semiconductor junction, is still under discussion, (10) in particular when moving away from the nanoparticle picture toward more complex nanostructures or multilayer systems. (11) Designing novel devices and evaluating their performance therefore remain key to progress toward functional and sustainable solutions for advanced wastewater treatment.
Here, we present a combination of a periodically nanostructured broadband light-trapping three-dimensional surface, plasmonic metals like gold and aluminum (Au and Al), and a thin TiO2 layer deposited by atomic layer deposition (ALD). Our approach avoids nanostructuring of noble metals, which is typically achieved by complex and time-consuming wet chemical processing. (12) Furthermore, similar to TiO2 nanoparticles, the impact of such metal nanoparticles on living organisms and the environment remains unclear. (13−15) We circumvent these issues by incorporating the nanostructures in a glass (SiO2) substrate and then creating different material stacks with high surface areas on top of it. We demonstrate the efficiency of our photocatalytic devices through the reduction of methylene blue (MB) dye under different illumination wavelengths. Moreover, we show the successful degradation of the pharmaceutical pollutant carbamazepine under UV-A light using a nanostructured Au surface. Finally, we discuss our results with respect to the state of the art.

Results and Discussion

ARTICLE SECTIONS
Jump To

We carried out photocatalysis experiments on five different sample types to evaluate the influence of the nanostructures alone versus their combination with the plasmonic metals (Figure 1a, for process details see Methods section). The first reference sample (labeled “SiO2” in figures) was a bare fused silica wafer cut to the standard experiment size, which was assessed to determine the effects of photolysis (if present). The flat surface sample with 25 nm TiO2 was included again for investigating not only photolysis but also standard photocatalysis of the TiO2 itself (labeled “TiO2”). The third reference sample, 25 nm TiO2 on nanostructured SiO2, is expected to utilize the effects of light trapping and surface area increase (labeled “cones + TiO2”). The last two samples differ in the choice of metal (Al and Au) and comprise identically nanostructured SiO2 covered with each metal, followed by the 25 nm TiO2 ALD layer (labeled “cones + Al + TiO2” and “cones + Au + TiO2”). These samples combine the light trapping and surface area increase with the plasmonic effects in the metals and the Schottky barrier/electron transfer effects from the metal–semiconductor junction.

Figure 1

Figure 1. (a) Schematic of the final structures with active layer material combinations under investigation. (b) AFM image of the Al nanodisc hard mask created by nanoimprint lithography. (c) 3D AFM image of the conical nanostructure etched in SiO2. (d) Photograph of the Au/TiO2 photocatalytic panel. The colorful surface effect is observed under ambient light at an angle by the naked eye or with a camera and is due to diffraction and thin film interference effects.

Optical Absorption and Spectral Overlap Model

The photocatalytic activity of the samples is expected to correlate with their optical absorption. Hence, the optical transmittance and reflectance spectra of the samples were measured by ultraviolet to visible light (UV–vis) spectrophotometry (PerkinElmer Lambda 1050) in the wavelength range from 250 to 800 nm. The samples were laid flat against the ports of a 150 mm integrating sphere housing the photodetector (photomultiplier tube). This ensured the collection of total transmittance (T) and total reflectance (R) for each sample, from which we calculated the total absorbance A = 1 – TR.
The UV–vis spectra in Figure 2a show that the flat TiO2 reference has an absorption onset at 348 nm (dashed black line). It absorbs 55% of the UV-C radiation at 254 nm. The second reference sample with cones and TiO2 (black line) shows a significant enhancement of UV absorption by 20% and an expansion of the absorption onset to 392 nm, which can be attributed to the light-trapping cones. There is an additional absorption peak of 26% at ∼500 nm, matching the period of the structure. This is because the grid acts as a diffraction grating for light at a wavelength comparable to its period. Here, the subwavelength structure for light with wavelengths larger than the period is turned into a partially diffracting grating for wavelengths smaller than the period. These are known optical effects of gratings, which are less relevant for the photocatalytic application. (16)

Figure 2

Figure 2. (a) Measured optical absorbance of the four fabricated samples superimposed with UV-C and UV-A light source spectra. (b) FDTD simulations of the SiO2–Au–TiO2 nanocone structure: absorption cross-section in comparison to measured absorbance. (c) Simulated electric field cross-section (y-plane) at different illumination wavelengths with plasmonic field enhancement (“hot spot” formation).

The samples with Al (teal) and Au (orange) layers both show 95–97% absorption in the UV (250–320 nm) and over 70% across the visible and NIR spectrum (400–800 nm). Both types of metal/TiO2 samples have low reflectance over the visible wavelength range, on average 11% for the Al sample and 3% for the Au sample (Figure S3 in the Supporting Information). In the case of gold, absorption reaches a maximum of 97% at 500 nm; again, this effect is attributed to the grating period of the nanostructure. The Au-sample absorption spectrum is reproducible across different positions on the sample surface and different 5 × 5 cm2 panels (Figure S2). Our design was thus successful in creating highly efficient photocatalytic surfaces extended to a broader wavelength range. Figure 2a further shows the source spectra of the UV-C and UV-A lamps we use in the reactor setup, with their peaks at 254 and 365 nm, respectively. This visualizes the spectral overlap between the light source driving the reaction and the absorption capabilities of the photocatalytic surfaces.
The photocatalytic efficiency can be estimated with an optical overlap model for plasmon resonance energy transfer (PRET), (17) which relates and quantifies the photocatalytic reaction with the electric field induced by surface plasmon resonance (SPR). According to the model, the photoactivity of a reaction is directly proportional to the overlap between the illumination source spectrum, the semiconductor absorption spectrum, and the metal SPR spectrum. Typically, nanometer-sized edges, holes, or valleys in a metal surface can give rise to plasmonic “hot spots” or very intense electric fields in the vicinity of the metallic nanostructure. The energy is transferred non-radiatively from the plasmon to the semiconductor photocatalyst via the electric field in the near-field zone, generating electron–hole pairs in the photocatalyst.
We have simulated the optical properties of a periodic array of SiO2–Au–TiO2 nanocones on a SiO2 substrate via finite-difference time domain (FDTD) solver of Maxwell’s equations by Lumerical (for details, see Methods section). From this simulation, we have obtained the absorption cross-section for incident wavelengths between 200 and 800 nm, as well as the e electric field distribution in the xz plane at different incident wavelengths. Figure 2b shows the simulated absorption cross-section together with the measured absorbance and distinguishes between three spectral regimes. Namely, starting from the shortest wavelengths, the TiO2, Al, and Au absorption regions extend to ∼380 nm, where all three materials intensively absorb incoming photons. Then, between ∼380 and ∼520 nm, only Au still has significant interband absorption, while the cone geometry still traps incoming light. Last, above ∼520 nm and up until 800 nm, the dielectric properties of Au (the negative real part of the dielectric function and the imaginary part close to zero, see Figure S7) can enable plasmonic field enhancement.
In addition, Figure 2c shows cross-sectional images of the simulated electric field at illumination wavelengths from this last spectral region above 520 nm. This is where we expect to see the electric field hot spots. The simulations show that they are indeed present, localized on the cone tip for shorter wavelengths of 560 to 590 nm, then at the sidewalls for 662 to 705 nm. As the wavelengths are getting larger than the surface topography dimensions, the simulation shows the electric field building up between the sidewalls of neighboring cones and also in the trenches (753 nm).
Based on the spectral overlap between sample absorption and lamp irradiation, we predict the highest photocatalytic efficiency for the Au sample at UV-C and at UV-A and possibly also for visible wavelengths due to the stronger extinction peak at 520 nm compared to the Al sample. The flat sample is expected to show the lowest efficiency.

Methylene Blue Degradation

We have performed dye degradation experiments under different illumination regimes corresponding to different regions of the samples’ optical spectra in order to distinguish between the effects of photolysis, standard semiconductor photocatalysis, and the enhanced photocatalysis of the combined metallic and TiO2 surfaces. The theoretical predictions based on the overlap model were experimentally evaluated using a self-built small-scale reactor in continuous flow mode (Figure 3a,b).

Figure 3

Figure 3. (a) Photocatalytic test reactor with lamp mount. Inset: basin with a 5 × 5 cm2 sample and MB aqueous solution (0.1 L) under illumination. (b) 3D sketch of the reactor, comprising light source over the testing area and water pump with piping to recirculate the liquid. (c) Close-up of the sensor system with red LED and photodiode (PD). The blue arrow indicates the direction of liquid flow through the well-defined volume of the sensor body.

The sample under study lies in a basin, while an aqueous solution recirculates over it and a lamp shines light on it for the duration of the experiment. We chose three different lamps to study different regimes of the TiO2 photocatalytic reaction: in-bandgap (UV-C) and out-of-bandgap (narrow band UV-A and broadband white light). The experiments with methylene blue (MB) were started with 30 min in the dark to allow for adsorption of MB on the sample surface. The MB degradation was monitored in real time by recording the reduction of its absorption peak at 663 nm with a red LED and a photodiode-based sensor (Figure 3c). Every 10 s, the red LED shines through a well-defined volume of the MB solution and the photodiode measures the transmitted light. Hence, the photodiode signal is a measure of the water transparency, which corresponds to a decrease in MB concentration C/C0, where C is the current concentration and C0 is the initial concentration. The initial concentration of the MB solution was 10–5 mol L–1, and the volume of the solution was 0.1 L. The transparency at this concentration level was measured and normalized to 1 in the figures, while a water transparency of 0 corresponds to clear water according to the detection limit of the sensor. The sensor was calibrated such that the light absorption intensity has a linear dependence on the MB concentration for the measurement range (see Figure S5). The detection range was tuned with the sensor electronics for a large sensitivity between the initial and the final concentration. The water transparency level, related to C/C0, should therefore always lie between one and zero. However, recordings of values larger than one are possible due to water evaporation and do not indicate an increase in concentration beyond the initial C0.
First, we investigated the UV-C regime with a lamp peak at λ = 254 nm. The corresponding photon energies are larger than the bandgap of TiO2, and there is high intrinsic absorption by the semiconductor. The real-time increase of water transparency under UV-C for all samples is shown in Figure 4a. The MB degradation measured on the reference-fused silica substrate reflects photolysis (dashed blue line). After 3 h of runtime, the final water transparency level due to photolysis only was 0.86. Similar results were obtained with the flat TiO2 ALD layer (dashed black line) with water transparency reaching 0.84, indicating that the photocatalytic activity of the as-deposited flat thin film is negligible compared to the pure photolysis under UV-C. In contrast, the performance improved to a relative transparency level of 0.76 when using the sample with TiO2 on the cone surface (solid black line in Figure 4a). The reactor run with this sample combines the effects of photolysis and standard photocatalysis due to TiO2 after the surface area enlargement. The increase in MB degradation is caused by the larger surface that can take part in the exchange of carriers with adsorbed dye molecules. This measurement then is a reference for the metallic samples, since structuring and layer thicknesses remain the same. As soon as one of the two metals are integrated on top of the glass nanostructure and are covered by the same TiO2 layer, other enhancing effects come into play. The data of the two metal samples in Figure 4a confirm the overlap model, as they show enhanced photocatalytic activity in UV-C compared to the TiO2 on cones reference. The water relative transparency reached 0.51 and 0.37 for Al (teal line) and Au (orange line), respectively.

Figure 4

Figure 4. Methylene blue degradation, reactor data recorded in real time. (a) Under UV-C (254 nm) illumination. Dashed blue line: photolysis measured with a blank SiO2 sample. Dashed black line: combined effect of photolysis and standard photocatalysis measured with a flat TiO2 layer. Solid black line: combined effect of photolysis and standard photocatalysis measured with TiO2 on a cone-structured surface. Enhanced photocatalysis is achieved by a stack of Al (teal line) or Au (orange line) and TiO2 on cone-structured surfaces. (b) Under UV-A (365 nm) illumination. No photolysis or standard photocatalysis present, as measured with both a flat TiO2 layer (dashed black line) and TiO2 layer on the cone-structured surface (solid black line). Photocatalysis driven by plasmonics is demonstrated by the Al/TiO2 (teal line) and Au/TiO2 (orange line) layer stacks on cone-structured surfaces. (c) First-order kinetics fit of samples under UV-A illumination. (d) Under white light from high-pressure lamp. No photocatalytic reaction with a flat TiO2 layer (dashed black line) or plasmonic photocatalyst with Al (teal line) present. Lines show an apparent increase in dye concentration because of the lamp’s excessive heating and subsequent water evaporation. The plasmonic photocatalyst with Au (orange line) successfully degrades the dye; the water transparency level increase due to evaporation is compensated for in this data set.

Second, we examined the UV-A regime with a lamp peak at λ = 365 nm, where there is no discernible photolysis and photon energies are below the bandgap of TiO2. Hence, we expect no photocatalytic dye degradation unless there is a plasmonic enhancement by the metallic cone structures. The water transparency kinetics during a 3 h run are shown in Figure 4b. The reference samples of flat TiO2 (dashed black line) and TiO2 on the cone surface (solid black line) remain at the initial MB concentration, which confirms the lack of photolysis and photocatalysis by TiO2 alone. The substrates with Al (teal line) and Au (orange line), in contrast, increased the water transparency to 0.82 and 0.57, respectively. This is evidence for a plasmonically enhanced photocatalytic reaction.
We have fitted the data for the UV-A regime to first-order kinetics in Figure 4c, using the calibration curve we had previously recorded (Figure S5). The apparent rate constants obtained from the slopes are kAu = 0.02 min–1 for the cones + Au + TiO2 sample and kAl = 0.005 min–1 for the cones + Al + TiO2. For the cones + TiO2 sample (no metal), there is no apparent photocatalytic reaction, the concentration stays very close to the initial value, and ln(C0/C) stays very close to zero. Therefore, the apparent negative slope with kTiO = −0.002 min–1 has no physical meaning. These results were later reproducibly confirmed on different days (example for the cones + Au + TiO2 sample on Figure S3).
Third, MB degradation was investigated under a broadband white light source. In this case, only the Au sample showed photocatalytic activity, reaching a final water transparency level of 0.48 after 3 h (Figure 4d). The measurements taken from the other samples show an apparent increase in dye concentration because the lamp’s excessive heat evaporated some of the water. This was observed despite a built-in Peltier cooling element, which was apparently insufficient. From Figure 4d and the calibration curve (Figure S5), we can estimate the influence of water evaporation to the dye concentration in the experiment with visible light. The final value of water transparency after 3 h of white light illumination for the TiO2 sample (no photocatalysis present) is 1.07, which is equivalent to a 1.8 V sensor signal. Given that at 2 V, the MB concentration is 1 × 10–5 mol L–1 and the linear fit of our calibration curve, we estimate that 1.8 V corresponds to a 1.13 × 10–5 mol L–1 MB concentration. This is an increase in concentration of only one-tenth of the initial value, and it leads to a water transparency decrease of 7% by the end of the measurement. Therefore, we consider the effect of water evaporation on the accuracy of the experiment to be within an acceptable margin, though modifications will be necessary for future experiments.
These three experiments serve to demonstrate the underlying mechanisms. In the UV-C case, the metallic nanostructures led to an improvement of efficiency by up to 51% at the 3 h mark for Au compared to conventional photocatalysis and photolysis. Here, the enhancement can be primarily attributed to the Schottky barrier at the metal–semiconductor interface, which lowers recombination losses of carriers generated by absorption in the TiO2 layer. A linear extrapolation of the measurement data reveals that the metallic samples would clear the water volume completely 2.2 times (Al) to 2.8 times (Au) faster than the TiO2 alone. In the UV-A case, the TiO2 alone barely shows a photocatalytic reaction throughout the measured test interval. Here, the metallic nanostructures are required to induce photocatalytic dye degradation. In the white light case, Au nanostructures are required to induce catalysis.
The superior performance of the Au sample compared to the Al sample at UV-C cannot be conclusively attributed to a difference in optical absorption at this wavelength (compare Figure 2a). This indicates that other enhancement effects, like the Schottky barrier at the metal–semiconductor junction, are driving the improved reaction in the in-gap regime. A higher Schottky barrier would promote better separation of the photoexcited charge carriers. Assuming the same electron affinity of TiO2 in both cases, the Schottky barrier will be higher for the metal with the larger work function, as the barrier is given by the difference between the work function of the metal and the electron affinity of the semiconductor. (17) The work function of Au is typically in the range of 5.31–5.47 eV, while for Al it is 4.06–4.41 eV. (18) Therefore, the Au–TiO2 interface would have the higher potential barrier and would be better at preventing electron–hole recombination.
The Schottky barrier plays an important role for enhancing photocatalytic performance by reducing recombination losses, if the attached semiconductor generates electron–hole pairs due to absorption. The curves measured under white light exposure (Figure 4d) nonetheless show that neither the TiO2 nor the Al–TiO2 produces a decent measurable dye degradation. At wavelengths in the out-of-gap regime, the semiconductor does not absorb light and without generated carriers, a sole reduction of recombination losses clearly cannot lead to any enhancement. Still, the Au–TiO2 sample shows ample dye degradation. The carriers for this reaction presumably result from the interband absorption of light in the Au layer, followed by carrier transfer to the semiconductor. This transfer is most likely caused by a plasmonic enhancement in the nanopatterned metal. Here, the plasmonic field enhancement generates hot carriers in the metal that may overcome the Schottky barrier and be transferred to the semiconductor. The field enhancement is depicted in the simulation of the absorption cross-sections in Figure 2b. It is strongest between 620 and 750 nm where the imaginary part of the dielectric function of Au is lowest (Figure S7). The E-field enhancement is visualized as hot spots in the structure cross-sections in Figure 2c derived from the simulation data.
Benchmarking the performance of our photocatalytic surfaces to the state of the art in literature proves quite difficult because there is great variation in experimental parameters and conditions across reports─from sample composition, to irradiation wavelength and power, to test substance type, volume, and concentration. Pedanekar et al. (19) recently published a table that collects the percentage of degradation efficiency in time as reported by different researchers. We summarize here only entries with MB as the model pollutant (Table 1) and conclude that our best-performing sample (60% in 180 min) matches well the state of the art for thin film photocatalysis. On closer inspection of the references, it becomes apparent that the cited percentages and times mostly refer to smaller testing volumes, for instance: 0.0035–0.005 (20,21) or 0.03, (22) 0.035, (23) 0.05 L, (24) compared to 0.1 L in our case. The initial MB concentration of 10–5 mol L–1 in the literature references is the same as in our experiments. The only other reference with a testing volume of 0.1 L reports 60% degradation in 300 min. (25) Therefore, we conclude that our photocatalytic surfaces indeed constitute the state of the art in efficiency of purifying larger quantities of water.
Table 1. Photocatalytic Performance of Various Metal Oxide and Sulfide Thin Filmsa
thin film photocatalystsynthesis methodirradiationtesting volume [L]b/sample area [cm2]percentage of degradation efficiency in timerefs
N-doped TiO2sol–gelvisible (150 W, Xe > 400 nm)0.005, 4.889% in 150 min (20)
Cu-doped TiO2/reduced graphene oxidespray pyrolysisUV-A, UV-B (300 W, 280–400 nm)0.035, 463% in 180 min (23)
ZnO-doped SiO2sol–gelUV-A (20 W)0.1, 160% in 300 min (25)
ZnSchemical bath depositionUV-C (252 nm, 11 W)0.0035, 292% in 240 min (21)
Bi2VO5.5/Bi2O3chemical solution depositionUV (300 W, Xe)0.03, 1689.97%. in 300 min (22)
CdSSILARUV–visible (500 W, Xe)0.05 powder89% in 120 min (24)
TiO2–Aumagnetron sputtering/ALDUV-A (365 nm, 6 W), visible (125 W)0.1, 2560% in 180 min, 50% in 180 minthis work
a

Adapted with permission from ref (19). Copyright 2020 Elsevier.

b

All entries are for methylene blue with concentration of 10–5 mol/L.

We further benchmarked our results with similar plasmonic photocatalysts and plotted the degradation time to reach maximum transparency (apparent C/C0 = 0) in Figure 5a. We included several reports based on sputtered TiO2 films decorated with metal nanoparticles. (4,26,27) We also added results from metal-doped TiO2 nanoparticle films fabricated by sol–gel methods (28−31) and TiO2 layers without metals. (32) In the interest of fair comparison and to illustrate the complexity of making fair comparisons between vastly different approaches to photocatalysis and diversity in results reporting and presentation, we have also included (Figure 5b) a table with the values of the apparent rate constant k reported in the studies referenced in Table 1 and Figure 5a. Although the shortest times and highest values of k were achieved with TiO2/metallic nanoparticles, there is still the open question of environmental exposure of nanoparticles. In addition, our data is in the same range as these nanoparticle-based methods and provides stable TiO2 thin film surfaces and may therefore be preferable overall. Looking beyond this initial performance evaluation, the possibility of iterative optimization of the nanostructure parameters and the choice of materials, together with the upscaling potential through large-area nanoimprint, makes the proposed photocatalytic surfaces an attractive concept for further studies.

Figure 5

Figure 5. Comparison of typical MB degradation speed for different TiO2-based photocatalysts. (a) In terms of time for C/C0 to reach apparent zero. Literature sources are indicated with references next to the points. Insets: SEM images of (i) gold-covered conical nanostructures as fabricated by us, (ii) a flat TiO2 ALD layer as deposited by us on silicon. (b) In terms of apparent rate constant k for different references.

Carbamazepine Degradation

As a realistic real-life scenario, we investigated the degradation of the pollutant carbamazepine under UV-A illumination choosing the Au sample for its superior demonstrated performance. The sample surface, the reactor body, and the tubing were thoroughly cleaned with distilled water to remove MB residues from previous experiments. Otherwise, the experimental system and the photocatalytic sample remained unmodified. However, carbamazepine has no absorption peaks in the visible spectrum. This prevented the real-time monitoring in the reactor with the red LED and photodiode combination. Instead, we measured the optical absorption spectra of the initial carbamazepine solution and the solution after 6 h of processing inside the reactor via our UV–vis spectrophotometer with 1 cm path length quartz cuvettes. The total solution volume in the experiment was 0.1 L, and the initial carbamazepine concentration was 1 mg/L. For reference, carbamazepine detected in wastewater samples and in environmental waters is typically in the ng/L to μg/L range. (33) After 6 h of reactor runtime, the characteristic carbamazepine absorption peak at 284 nm falls from 28.4 to 14.7% (Figure 6). The second typical absorption peak close to 200 nm lies at the spectral limit of the measurement tool; therefore, it is much harder to quantify. Nevertheless, the graph shows a clear reduction of the optical absorption─and hence of the concentration─of carbamazepine in the water, demonstrating the viability of our plasmonic nanostructure concept.

Figure 6

Figure 6. Carbamazepine degradation under UV-A illumination. Solution absorbance measured in UV–vis spectrophotometer before and after reactor run time. Black line: absorbance spectrum of the solution with an initial concentration of 1 mg/L. Orange line: absorbance spectrum of the solution after 6 h in the reactor with Au plasmonic photocatalyst. Characteristic peak at 284 nm reduced by 52%.

Conclusions

ARTICLE SECTIONS
Jump To

We have demonstrated enhanced photocatalysis efficiency via nanostructured and plasmonically engineered surfaces that act in synergy to circumvent the known limitations of thin film catalysts. Nanostructuring provides a three-fold surface area enlargement to increase the adsorption of reactants. In addition, the subwavelength dimensions of conical structures create a trapping effect for UV and visible light. This red-shifts the TiO2 absorption onset to the violet range of visible light and enhances the UV absorption maximum to 78.4% for a TiO2 thin film of merely 25 nm. Finally, a metal–semiconductor heterostructure formed on top of the textured surfaces (Al or Au and a TiO2 ALD thin layer) further enhances the photocatalytic activity due to both promotion of charge carrier separation and plasmonic resonance. Specifically, the addition of Al or Au accelerates organic dye degradation under UV-C irradiation by a factor of 2.8 compared to bare TiO2 and activates the photocatalytic degradation under UV-A, achieving up to 43% degradation in 180 min for methylene blue and 48% for carbamazepine in 360 min. Furthermore, the large plasmon resonance peak of the Au sample around 520 nm leads to methylene blue degradation of 52% in 180 min under broadband white light, which is relevant for solar-driven applications. We have achieved these results with substrates that were manufactured with scalable nanostructuring and conventional thin-film deposition methods on standard SiO2 substrates. Furthermore, the approach is suitable for addressing other applications of photocatalysis, while being potentially less toxic than commonly employed nanoparticles. Our results hold promise for further up-scaling of thin film-based surfaces for wastewater treatment and other environmentally relevant sectors.

Methods

ARTICLE SECTIONS
Jump To

We first modeled the devices optical response via finite-difference time domain (FDTD) simulations with Lumerical FDTD photonic simulation software to predict a suitable nanostructure grating periodicity. The critical features for the gratings are their period/distance between structures and their depth of texturing (Figure 1a). One prerequisite is that the grating period needs to be smaller than the wavelength of the incident light. Here, a period of less than or equal to ∼500 nm is required for the enhancement of optical absorption in the UV and visible range. In addition, light-trapping is achieved when the height of the structures is equal to or larger than the wavelength of light, that is, more than 300 nm in the case of UV-solar irradiation.
The chosen geometry closely matched the fabricated structure shape, including the dimensions of period and a height of 480 nm, as well as thin film thicknesses of ∼20 nm. Periodic boundary conditions were applied in the x and y direction, while a perfectly matched layer condition was set in the z direction. The simulation did not take into account surface roughness and effects arising from imperfections in the fabricated arrays. It was also limited to showing the response of the structures when the external electromagnetic field of the light interacts with them, which is not enough to explain the full photocatalytic reaction. Still, the simulation is a good first approximation to guide our understanding of the experimental results.
Common semiconductor technology tools and techniques can successfully produce structures that meet the theoretical requirements. Although this fabrication route may at first seem unusual in the context of TiO2 photocatalysis, it has been a viable choice for other solar-driven devices like nanostructured solar cells. (34) Therefore, employing it for photocatalytic surfaces has the potential to enable innovations within the existing industrial infrastructure, as well as fabricating larger panels to cover the large surface areas required for wastewater treatment plants.
We used nanoimprint lithography (NIL) to fabricate nanostructures on 150 mm SiO2 wafers (fused silica), which provides control of structural parameters like periodicity and lateral dimensions on the order of 10 nm. (35) From a single mold or master wafer, tens of nanostructured stamps can be readily produced and each of them can be used to replicate up to 60 nanostructured wafers, corresponding to an area of ∼1 m2. Here, the master was made with laser interference lithography (LIL) in a setup operating with a laser wavelength of 266 nm. We exposed a 150 mm silicon wafer covered with UV resist with two expanded laser beams that interfered on the substrate (Figure S1 in Supporting Information, process details in ref (36)). This created a periodic 2D grid of light and dark spots in the resist. The substrate was then rotated by 90°, and a second exposure produced a pillar grating with pillar diameters of 240 nm, equal to half of the period defined as the distance from one pillar center to the center of an adjacent pillar (480 nm). Next, the resist grating was transferred into the silicon by reactive ion etching, and the patterned silicon surface was coated with an anti-adhesion layer of octafluorocyclobutane. We then made a reusable soft stamp by casting polydimethylsiloxane (PDMS) onto the imprint master wafer.
The PDMS soft stamp was used to lithographically define the nanostructures of the fused silica wafers. However, the actual etching of the 450 nm deep conical structures in these wafers required an additional Al hard mask, which was deposited by DC magnetron sputtering. Next, AMONIL resist (37) was spin-coated, the soft stamp was pressed onto the resist, the resist was cured under UV illumination, and the stamp was removed. Afterward, reactive ion etching with BCl3-based plasma was used to transfer the pattern from the resist into the Al layer, forming Al nanodiscs as a hard mask on the SiO2 surface (Figure 1b). A second etching step with CHF3-based plasma produced cones in the SiO2 with an average height of 450 nm (Figure 1c). The wafers were diced into 5 × 5 cm2-sized panels for further processing.
Either one of the two metals (Au, Al) was deposited onto the cones via DC magnetron sputtering. The nominal thickness of the metals was 40–50 nm, that is, if they had been deposited on a flat surface. The actual thickness over the conical topography can be roughly estimated by assuming a uniform distribution of the metal over the cones (SEM image in Figure 5a). The geometry of the samples results in a threefold increase of the surface area compared to a flat surface. Thus, the average metal thickness is approximately 15 nm.
Finally, a TiO2 layer with a thickness of tTiO2 = 25 nm was deposited on top of the metals with a conformal ALD plasma process based on a titanium tetrachloride (TiCl4) precursor at 300 °C. Figure 1d is a photograph of the panel surface after nanostructuring and Au/TiO2 thin-film deposition. The structured glass acts as a surface relief diffraction grating for the incoming ambient light. This effect, combined with the thin film interference effect, results in a colorful pattern observable at an angle by the naked eye or with a camera. In addition to the Au/TiO2 and Al/TiO2 panels, a flat TiO2 film on plain SiO2 and a cone sample without metal but with TiO2 were also fabricated as references, both with tTiO2 = 25 nm, in addition to a reference fused silica wafer.
For the MB test in a laboratory-scale continuous flow photocatalytic reactor, our setup design and calibration were guided by the standard ISO 10678:2010 (38) and consulting with the suggestions of Prof. Andrew Mills in his review of the ISO standard. (39)

Supporting Information

ARTICLE SECTIONS
Jump To

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsanm.3c02867.

  • Additional experimental details, including: a schematic of the laser interference lithography setup used to fabricate master wafers for nanoimprint lithography; optical absorbance spectra of the nanostructured plasmonic photocatalyst with Au and TiO2 measured on two different panels at three different positions each; and degradation of methylene blue by the nanostructured plasmonic photocatalyst with Au and TiO2 under UV-A with reactor data from measurements on different days (PDF)

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

ARTICLE SECTIONS
Jump To

  • Corresponding Author
  • Authors
    • Desislava Daskalova - Advanced Microelectronic Center Aachen, AMO GmbH, 52074 Aachen, GermanyChair of Electronic Devices, RWTH Aachen University, 52074 Aachen, GermanyOrcidhttps://orcid.org/0000-0002-7636-413X
    • Gonzalo Aguila Flores - Advanced Microelectronic Center Aachen, AMO GmbH, 52074 Aachen, Germany
    • Ulrich Plachetka - Advanced Microelectronic Center Aachen, AMO GmbH, 52074 Aachen, Germany
    • Michael Möller - Advanced Microelectronic Center Aachen, AMO GmbH, 52074 Aachen, Germany
    • Julia Wolters - Institute of Environmental Engineering, RWTH Aachen University, 52074 Aachen, Germany
    • Thomas Wintgens - Institute of Environmental Engineering, RWTH Aachen University, 52074 Aachen, Germany
  • Notes
    The authors declare no competing financial interest.

Acknowledgments

ARTICLE SECTIONS
Jump To

We acknowledge funding through the German Ministry of Education and Research, BMBF in the project PEPcat (02WCL 1519A) and the European Union’s Horizon Europe research and innovation program under the grant agreement No 101084261 (FreeHydroCells).

References

ARTICLE SECTIONS
Jump To

This article references 39 other publications.

  1. 1
    Zhang, H.; Chen, G.; Bahnemann, D. W. Photoelectrocatalytic Materials for Environmental Applications. J. Mater. Chem. 2009, 19, 50895121,  DOI: 10.1039/B821991E
  2. 2
    Pan, J. H.; Dou, H.; Xiong, Z.; Xu, C.; Ma, J.; Zhao, X. S. Porous Photocatalysts for Advanced Water Purifications. J. Mater. Chem. 2010, 20, 45124528,  DOI: 10.1039/B925523K
  3. 3
    Colvin, V. L. The Potential Environmental Impact of Engineered Nanomaterials. Nat. Biotechnol. 2003, 21, 11661170,  DOI: 10.1038/nbt875
  4. 4
    Vahl, A.; Veziroglu, S.; Henkel, B.; Strunskus, T.; Polonskyi, O.; Aktas, O. C.; Faupel, F. Pathways to Tailor Photocatalytic Performance of TiO2 Thin Films Deposited by Reactive Magnetron Sputtering. Materials 2019, 12, 2840,  DOI: 10.3390/ma12172840
  5. 5
    Cheng, H.-E.; Hung, C.-H.; Yu, I.-S.; Yang, Z.-P. Strongly Enhancing Photocatalytic Activity of TiO2 Thin Films by Multi-Heterojunction Technique. Catalysts 2018, 8, 440,  DOI: 10.3390/catal8100440
  6. 6
    Dharma, H. N. C.; Jaafar, J.; Widiastuti, N.; Matsuyama, H.; Rajabsadeh, S.; Othman, M. H. D.; Rahman, M. A.; Jafri, N. N. M.; Suhaimin, N. S.; Nasir, A. M.; Alias, N. H. A Review of Titanium Dioxide (TiO2)-Based Photocatalyst for Oilfield-Produced Water Treatment. Membranes 2022, 12, 345,  DOI: 10.3390/membranes12030345
  7. 7
    Dasineh Khiavi, N.; Katal, R.; Kholghi Eshkalak, S.; Masudy-Panah, S.; Ramakrishna, S.; Jiangyong, H. Visible Light Driven Heterojunction Photocatalyst of CuO–Cu2O Thin Films for Photocatalytic Degradation of Organic Pollutants. Nanomaterials 2019, 9, 1011,  DOI: 10.3390/nano9071011
  8. 8
    Zheng, Z.; Xie, W.; Huang, B.; Dai, Y. Plasmon-Enhanced Solar Water Splitting on Metal-Semiconductor Photocatalysts. Chem. - Eur. J. 2018, 24, 1832218333,  DOI: 10.1002/chem.201803705
  9. 9
    Kowalska, E. Plasmonic Photocatalysts. Catalysts 2021, 11, 410,  DOI: 10.3390/catal11040410
  10. 10
    Abed, J.; Rajput, N. S.; Moutaouakil, A. E.; Jouiad, M. Recent Advances in the Design of Plasmonic Au/TiO2 Nanostructures for Enhanced Photocatalytic Water Splitting. Nanomaterials 2020, 10, 2260,  DOI: 10.3390/nano10112260
  11. 11
    Zhang, P.; Wang, T.; Gong, J. Mechanistic Understanding of the Plasmonic Enhancement for Solar Water Splitting. Adv. Mater. 2015, 27, 53285342,  DOI: 10.1002/adma.201500888
  12. 12
    Dimcheva, N. Nanostructures of Noble Metals as Functional Materials in Biosensors. Curr. Opin. Electrochem. 2020, 19, 3541,  DOI: 10.1016/j.coelec.2019.09.008
  13. 13
    Siegel, J.; Staszek, M.; Polívková, M.; Valová, M.; Šuláková, P.; Švorčík, V. Structure-Dependent Biological Response of Noble Metals: From Nanoparticles, through Nanowires to Nanolayers; IntechOpen, 2017.
  14. 14
    Carnovale, C.; Bryant, G.; Shukla, R.; Bansal, V. Identifying Trends in Gold Nanoparticle Toxicity and Uptake: Size, Shape, Capping Ligand, and Biological Corona. ACS Omega 2019, 4, 242256,  DOI: 10.1021/acsomega.8b03227
  15. 15
    Sani, A.; Cao, C.; Cui, D. Toxicity of Gold Nanoparticles (AuNPs): A Review. Biochem. Biophys. Rep. 2021, 26, 100991,  DOI: 10.1016/j.bbrep.2021.100991
  16. 16
    Cheben, P.; Halir, R.; Schmid, J. H.; Atwater, H. A.; Smith, D. R. Subwavelength Integrated Photonics. Nature 2018, 560, 565572,  DOI: 10.1038/s41586-018-0421-7
  17. 17
    Khan, M. R.; Chuan, T. W.; Yousuf, A.; Chowdhury, M. N. K.; Cheng, C. K. Schottky Barrier and Surface Plasmonic Resonance Phenomena towards the Photocatalytic Reaction: Study of Their Mechanisms to Enhance Photocatalytic Activity. Catal. Sci. Technol. 2015, 5, 25222531,  DOI: 10.1039/C4CY01545B
  18. 18
    Hölzl, J.; Schulte, F. K. Work Function of Metals. In Solid Surface Physics; Hölzl, J., Schulte, F. K., Wagner, H., Eds.; Springer Tracts in Modern Physics; Springer: Berlin, Heidelberg, 1979; pp 1150.
  19. 19
    Pedanekar, R. S.; Shaikh, S. K.; Rajpure, K. Y. Thin Film Photocatalysis for Environmental Remediation: A Status Review. Curr. Appl. Phys. 2020, 20, 931952,  DOI: 10.1016/j.cap.2020.04.006
  20. 20
    Lee, H.-K.; Fujiwara, T.; Okada, T.; Fukushima, T.; Lee, S.-W. Fabrication of Visible-Light Responsive N-Doped TiO2 Nanothin Films via a Top–down Sol–Gel Deposition Method Using NH4TiOF3 Single Crystals. Chem. Lett. 2018, 47, 628631,  DOI: 10.1246/cl.180005
  21. 21
    Cruz, J. S.; Cruz, D. S.; Arenas-Arrocena, M. C.; De, F.; Flores, M.; Hernández, S. A. M. Green Synthesis of ZnS Thin Films by Chemical Bath Deposition. Chalcogenide Lett. 2015, 12, 277285
  22. 22
    Xie, W.; Zhong, L.; Wang, Z.; Liang, F.; Tang, X.; Zou, C.; Liu, G. Photocatalytic Performance of Bi2VO5.5/Bi2O3 Laminated Composite Films under Simulated Sunlight Irradiation. Solid State Sci. 2019, 94, 17,  DOI: 10.1016/j.solidstatesciences.2019.05.010
  23. 23
    Pham, T.-T.; Nguyen-Huy, C.; Lee, H.-J.; Nguyen-Phan, T.-D.; Son, T. H.; Kim, C.-K.; Shin, E. W. Cu-Doped TiO2/Reduced Graphene Oxide Thin-Film Photocatalysts: Effect of Cu Content upon Methylene Blue Removal in Water. Ceram. Int. 2015, 41, 1118411193,  DOI: 10.1016/j.ceramint.2015.05.068
  24. 24
    Ravichandran, K.; Porkodi, S. Addressing the Issue of Under-Utilization of Precursor Material in SILAR Process: Simultaneous Preparation of CdS in Two Different Forms–Thin Film and Powder. Mater. Sci. Semicond. Process. 2018, 81, 3037,  DOI: 10.1016/j.mssp.2018.02.037
  25. 25
    Ali, A. M.; Ismail, A. A.; Najmy, R.; Al-Hajry, A. Preparation and Characterization of ZnO–SiO2 Thin Films as Highly Efficient Photocatalyst. J. Photochem. Photobiol., A 2014, 275, 3746,  DOI: 10.1016/j.jphotochem.2013.11.002
  26. 26
    Ghori, M. Z.; Veziroglu, S.; Hinz, A.; Shurtleff, B. B.; Polonskyi, O.; Strunskus, T.; Adam, J.; Faupel, F.; Aktas, O. C. Role of UV Plasmonics in the Photocatalytic Performance of TiO2 Decorated with Aluminum Nanoparticles. ACS Appl. Nano Mater. 2018, 1, 37603764,  DOI: 10.1021/acsanm.8b00853
  27. 27
    Veziroglu, S.; Ghori, M. Z.; Obermann, A.-L.; Röder, K.; Polonskyi, O.; Strunskus, T.; Faupel, F.; Aktas, O. C. Ag Nanoparticles Decorated TiO2 Thin Films with Enhanced Photocatalytic Activity. Phys. Status Solidi A 2019, 216, 1800898,  DOI: 10.1002/pssa.201800898
  28. 28
    Kaleji, B. K.; Sarraf-Mamoory, R.; Fujishima, A. Influence of Nb Dopant on the Structural and Optical Properties of Nanocrystalline TiO2 Thin Films. Mater. Chem. Phys. 2012, 132, 210215,  DOI: 10.1016/j.matchemphys.2011.11.034
  29. 29
    Wada, N.; Yokomizo, Y.; Yogi, C.; Katayama, M.; Tanaka, A.; Kojima, K.; Inada, Y.; Ozutsumi, K. Effect of Adding Au Nanoparticles to TiO2 Films on Crystallization, Phase Transformation, and Photocatalysis. J. Mater. Res. 2018, 33, 467481,  DOI: 10.1557/jmr.2018.16
  30. 30
    Sonawane, R. S.; Kale, B. B.; Dongare, M. K. Preparation and Photo-Catalytic Activity of Fe:TiO2 Thin Films Prepared by Sol–Gel Dip Coating. Mater. Chem. Phys. 2004, 85, 5257,  DOI: 10.1016/j.matchemphys.2003.12.007
  31. 31
    Khairy, M.; Zakaria, W. Effect of Metal-Doping of TiO2 Nanoparticles on Their Photocatalytic Activities toward Removal of Organic Dyes. Egypt. J. Pet. 2014, 23, 419426,  DOI: 10.1016/j.ejpe.2014.09.010
  32. 32
    Wardhani, S.; Purwonugroho, D.; Fitri, C. W.; Prananto, Y. P. Effect of PH and Irradiation Time on TiO2-Chitosan Activity for Phenol Photo-Degradation. AIP Conf. Proc. 2018, 2021, 050009,  DOI: 10.1063/1.5062759
  33. 33
    Mestre, A. S.; Carvalho, A. P. Photocatalytic Degradation of Pharmaceuticals Carbamazepine, Diclofenac, and Sulfamethoxazole by Semiconductor and Carbon Materials: A Review. Molecules 2019, 24, 3702,  DOI: 10.3390/molecules24203702
  34. 34
    Beard, M. C.; Luther, J. M.; Nozik, A. J. The Promise and Challenge of Nanostructured Solar Cells. Nat. Nanotechnol. 2014, 9, 951954,  DOI: 10.1038/nnano.2014.292
  35. 35
    Koo, N.; Plachetka, U.; Otto, M.; Bolten, J.; Jeong, J.; Lee, E.; Kurz, H. The Fabrication of a Flexible Mold for High Resolution Soft Ultraviolet Nanoimprint Lithography. Nanotechnology 2008, 19, 225304,  DOI: 10.1088/0957-4484/19/22/225304
  36. 36
    Lemme, M. C.; Moormann, C.; Lerch, H.; Möller, M.; Vratzov, B.; Kurz, H. Triple-Gate Metal–Oxide–Semiconductor Field Effect Transistors Fabricated with Interference Lithography. Nanotechnology 2004, 15, S208S210,  DOI: 10.1088/0957-4484/15/4/016
  37. 37
    AMONIL. High performance UV nanoimprint resist; AMO GmbH. https://www.amo.de/products-services/amonil/ (accessed Dec 03, 2022).
  38. 38
    ISO 10678:2010; International Organization for Standardization. https://www.iso.org/standard/46019.html (accessed May 10, 2023).
  39. 39
    Mills, A. An Overview of the Methylene Blue ISO Test for Assessing the Activities of Photocatalytic Films. Appl. Catal., B 2012, 128, 144149,  DOI: 10.1016/j.apcatb.2012.01.019

Cited By

This article has not yet been cited by other publications.

  • Abstract

    Figure 1

    Figure 1. (a) Schematic of the final structures with active layer material combinations under investigation. (b) AFM image of the Al nanodisc hard mask created by nanoimprint lithography. (c) 3D AFM image of the conical nanostructure etched in SiO2. (d) Photograph of the Au/TiO2 photocatalytic panel. The colorful surface effect is observed under ambient light at an angle by the naked eye or with a camera and is due to diffraction and thin film interference effects.

    Figure 2

    Figure 2. (a) Measured optical absorbance of the four fabricated samples superimposed with UV-C and UV-A light source spectra. (b) FDTD simulations of the SiO2–Au–TiO2 nanocone structure: absorption cross-section in comparison to measured absorbance. (c) Simulated electric field cross-section (y-plane) at different illumination wavelengths with plasmonic field enhancement (“hot spot” formation).

    Figure 3

    Figure 3. (a) Photocatalytic test reactor with lamp mount. Inset: basin with a 5 × 5 cm2 sample and MB aqueous solution (0.1 L) under illumination. (b) 3D sketch of the reactor, comprising light source over the testing area and water pump with piping to recirculate the liquid. (c) Close-up of the sensor system with red LED and photodiode (PD). The blue arrow indicates the direction of liquid flow through the well-defined volume of the sensor body.

    Figure 4

    Figure 4. Methylene blue degradation, reactor data recorded in real time. (a) Under UV-C (254 nm) illumination. Dashed blue line: photolysis measured with a blank SiO2 sample. Dashed black line: combined effect of photolysis and standard photocatalysis measured with a flat TiO2 layer. Solid black line: combined effect of photolysis and standard photocatalysis measured with TiO2 on a cone-structured surface. Enhanced photocatalysis is achieved by a stack of Al (teal line) or Au (orange line) and TiO2 on cone-structured surfaces. (b) Under UV-A (365 nm) illumination. No photolysis or standard photocatalysis present, as measured with both a flat TiO2 layer (dashed black line) and TiO2 layer on the cone-structured surface (solid black line). Photocatalysis driven by plasmonics is demonstrated by the Al/TiO2 (teal line) and Au/TiO2 (orange line) layer stacks on cone-structured surfaces. (c) First-order kinetics fit of samples under UV-A illumination. (d) Under white light from high-pressure lamp. No photocatalytic reaction with a flat TiO2 layer (dashed black line) or plasmonic photocatalyst with Al (teal line) present. Lines show an apparent increase in dye concentration because of the lamp’s excessive heating and subsequent water evaporation. The plasmonic photocatalyst with Au (orange line) successfully degrades the dye; the water transparency level increase due to evaporation is compensated for in this data set.

    Figure 5

    Figure 5. Comparison of typical MB degradation speed for different TiO2-based photocatalysts. (a) In terms of time for C/C0 to reach apparent zero. Literature sources are indicated with references next to the points. Insets: SEM images of (i) gold-covered conical nanostructures as fabricated by us, (ii) a flat TiO2 ALD layer as deposited by us on silicon. (b) In terms of apparent rate constant k for different references.

    Figure 6

    Figure 6. Carbamazepine degradation under UV-A illumination. Solution absorbance measured in UV–vis spectrophotometer before and after reactor run time. Black line: absorbance spectrum of the solution with an initial concentration of 1 mg/L. Orange line: absorbance spectrum of the solution after 6 h in the reactor with Au plasmonic photocatalyst. Characteristic peak at 284 nm reduced by 52%.

  • References

    ARTICLE SECTIONS
    Jump To

    This article references 39 other publications.

    1. 1
      Zhang, H.; Chen, G.; Bahnemann, D. W. Photoelectrocatalytic Materials for Environmental Applications. J. Mater. Chem. 2009, 19, 50895121,  DOI: 10.1039/B821991E
    2. 2
      Pan, J. H.; Dou, H.; Xiong, Z.; Xu, C.; Ma, J.; Zhao, X. S. Porous Photocatalysts for Advanced Water Purifications. J. Mater. Chem. 2010, 20, 45124528,  DOI: 10.1039/B925523K
    3. 3
      Colvin, V. L. The Potential Environmental Impact of Engineered Nanomaterials. Nat. Biotechnol. 2003, 21, 11661170,  DOI: 10.1038/nbt875
    4. 4
      Vahl, A.; Veziroglu, S.; Henkel, B.; Strunskus, T.; Polonskyi, O.; Aktas, O. C.; Faupel, F. Pathways to Tailor Photocatalytic Performance of TiO2 Thin Films Deposited by Reactive Magnetron Sputtering. Materials 2019, 12, 2840,  DOI: 10.3390/ma12172840
    5. 5
      Cheng, H.-E.; Hung, C.-H.; Yu, I.-S.; Yang, Z.-P. Strongly Enhancing Photocatalytic Activity of TiO2 Thin Films by Multi-Heterojunction Technique. Catalysts 2018, 8, 440,  DOI: 10.3390/catal8100440
    6. 6
      Dharma, H. N. C.; Jaafar, J.; Widiastuti, N.; Matsuyama, H.; Rajabsadeh, S.; Othman, M. H. D.; Rahman, M. A.; Jafri, N. N. M.; Suhaimin, N. S.; Nasir, A. M.; Alias, N. H. A Review of Titanium Dioxide (TiO2)-Based Photocatalyst for Oilfield-Produced Water Treatment. Membranes 2022, 12, 345,  DOI: 10.3390/membranes12030345
    7. 7
      Dasineh Khiavi, N.; Katal, R.; Kholghi Eshkalak, S.; Masudy-Panah, S.; Ramakrishna, S.; Jiangyong, H. Visible Light Driven Heterojunction Photocatalyst of CuO–Cu2O Thin Films for Photocatalytic Degradation of Organic Pollutants. Nanomaterials 2019, 9, 1011,  DOI: 10.3390/nano9071011
    8. 8
      Zheng, Z.; Xie, W.; Huang, B.; Dai, Y. Plasmon-Enhanced Solar Water Splitting on Metal-Semiconductor Photocatalysts. Chem. - Eur. J. 2018, 24, 1832218333,  DOI: 10.1002/chem.201803705
    9. 9
      Kowalska, E. Plasmonic Photocatalysts. Catalysts 2021, 11, 410,  DOI: 10.3390/catal11040410
    10. 10
      Abed, J.; Rajput, N. S.; Moutaouakil, A. E.; Jouiad, M. Recent Advances in the Design of Plasmonic Au/TiO2 Nanostructures for Enhanced Photocatalytic Water Splitting. Nanomaterials 2020, 10, 2260,  DOI: 10.3390/nano10112260
    11. 11
      Zhang, P.; Wang, T.; Gong, J. Mechanistic Understanding of the Plasmonic Enhancement for Solar Water Splitting. Adv. Mater. 2015, 27, 53285342,  DOI: 10.1002/adma.201500888
    12. 12
      Dimcheva, N. Nanostructures of Noble Metals as Functional Materials in Biosensors. Curr. Opin. Electrochem. 2020, 19, 3541,  DOI: 10.1016/j.coelec.2019.09.008
    13. 13
      Siegel, J.; Staszek, M.; Polívková, M.; Valová, M.; Šuláková, P.; Švorčík, V. Structure-Dependent Biological Response of Noble Metals: From Nanoparticles, through Nanowires to Nanolayers; IntechOpen, 2017.
    14. 14
      Carnovale, C.; Bryant, G.; Shukla, R.; Bansal, V. Identifying Trends in Gold Nanoparticle Toxicity and Uptake: Size, Shape, Capping Ligand, and Biological Corona. ACS Omega 2019, 4, 242256,  DOI: 10.1021/acsomega.8b03227
    15. 15
      Sani, A.; Cao, C.; Cui, D. Toxicity of Gold Nanoparticles (AuNPs): A Review. Biochem. Biophys. Rep. 2021, 26, 100991,  DOI: 10.1016/j.bbrep.2021.100991
    16. 16
      Cheben, P.; Halir, R.; Schmid, J. H.; Atwater, H. A.; Smith, D. R. Subwavelength Integrated Photonics. Nature 2018, 560, 565572,  DOI: 10.1038/s41586-018-0421-7
    17. 17
      Khan, M. R.; Chuan, T. W.; Yousuf, A.; Chowdhury, M. N. K.; Cheng, C. K. Schottky Barrier and Surface Plasmonic Resonance Phenomena towards the Photocatalytic Reaction: Study of Their Mechanisms to Enhance Photocatalytic Activity. Catal. Sci. Technol. 2015, 5, 25222531,  DOI: 10.1039/C4CY01545B
    18. 18
      Hölzl, J.; Schulte, F. K. Work Function of Metals. In Solid Surface Physics; Hölzl, J., Schulte, F. K., Wagner, H., Eds.; Springer Tracts in Modern Physics; Springer: Berlin, Heidelberg, 1979; pp 1150.
    19. 19
      Pedanekar, R. S.; Shaikh, S. K.; Rajpure, K. Y. Thin Film Photocatalysis for Environmental Remediation: A Status Review. Curr. Appl. Phys. 2020, 20, 931952,  DOI: 10.1016/j.cap.2020.04.006
    20. 20
      Lee, H.-K.; Fujiwara, T.; Okada, T.; Fukushima, T.; Lee, S.-W. Fabrication of Visible-Light Responsive N-Doped TiO2 Nanothin Films via a Top–down Sol–Gel Deposition Method Using NH4TiOF3 Single Crystals. Chem. Lett. 2018, 47, 628631,  DOI: 10.1246/cl.180005
    21. 21
      Cruz, J. S.; Cruz, D. S.; Arenas-Arrocena, M. C.; De, F.; Flores, M.; Hernández, S. A. M. Green Synthesis of ZnS Thin Films by Chemical Bath Deposition. Chalcogenide Lett. 2015, 12, 277285
    22. 22
      Xie, W.; Zhong, L.; Wang, Z.; Liang, F.; Tang, X.; Zou, C.; Liu, G. Photocatalytic Performance of Bi2VO5.5/Bi2O3 Laminated Composite Films under Simulated Sunlight Irradiation. Solid State Sci. 2019, 94, 17,  DOI: 10.1016/j.solidstatesciences.2019.05.010
    23. 23
      Pham, T.-T.; Nguyen-Huy, C.; Lee, H.-J.; Nguyen-Phan, T.-D.; Son, T. H.; Kim, C.-K.; Shin, E. W. Cu-Doped TiO2/Reduced Graphene Oxide Thin-Film Photocatalysts: Effect of Cu Content upon Methylene Blue Removal in Water. Ceram. Int. 2015, 41, 1118411193,  DOI: 10.1016/j.ceramint.2015.05.068
    24. 24
      Ravichandran, K.; Porkodi, S. Addressing the Issue of Under-Utilization of Precursor Material in SILAR Process: Simultaneous Preparation of CdS in Two Different Forms–Thin Film and Powder. Mater. Sci. Semicond. Process. 2018, 81, 3037,  DOI: 10.1016/j.mssp.2018.02.037
    25. 25
      Ali, A. M.; Ismail, A. A.; Najmy, R.; Al-Hajry, A. Preparation and Characterization of ZnO–SiO2 Thin Films as Highly Efficient Photocatalyst. J. Photochem. Photobiol., A 2014, 275, 3746,  DOI: 10.1016/j.jphotochem.2013.11.002
    26. 26
      Ghori, M. Z.; Veziroglu, S.; Hinz, A.; Shurtleff, B. B.; Polonskyi, O.; Strunskus, T.; Adam, J.; Faupel, F.; Aktas, O. C. Role of UV Plasmonics in the Photocatalytic Performance of TiO2 Decorated with Aluminum Nanoparticles. ACS Appl. Nano Mater. 2018, 1, 37603764,  DOI: 10.1021/acsanm.8b00853
    27. 27
      Veziroglu, S.; Ghori, M. Z.; Obermann, A.-L.; Röder, K.; Polonskyi, O.; Strunskus, T.; Faupel, F.; Aktas, O. C. Ag Nanoparticles Decorated TiO2 Thin Films with Enhanced Photocatalytic Activity. Phys. Status Solidi A 2019, 216, 1800898,  DOI: 10.1002/pssa.201800898
    28. 28
      Kaleji, B. K.; Sarraf-Mamoory, R.; Fujishima, A. Influence of Nb Dopant on the Structural and Optical Properties of Nanocrystalline TiO2 Thin Films. Mater. Chem. Phys. 2012, 132, 210215,  DOI: 10.1016/j.matchemphys.2011.11.034
    29. 29
      Wada, N.; Yokomizo, Y.; Yogi, C.; Katayama, M.; Tanaka, A.; Kojima, K.; Inada, Y.; Ozutsumi, K. Effect of Adding Au Nanoparticles to TiO2 Films on Crystallization, Phase Transformation, and Photocatalysis. J. Mater. Res. 2018, 33, 467481,  DOI: 10.1557/jmr.2018.16
    30. 30
      Sonawane, R. S.; Kale, B. B.; Dongare, M. K. Preparation and Photo-Catalytic Activity of Fe:TiO2 Thin Films Prepared by Sol–Gel Dip Coating. Mater. Chem. Phys. 2004, 85, 5257,  DOI: 10.1016/j.matchemphys.2003.12.007
    31. 31
      Khairy, M.; Zakaria, W. Effect of Metal-Doping of TiO2 Nanoparticles on Their Photocatalytic Activities toward Removal of Organic Dyes. Egypt. J. Pet. 2014, 23, 419426,  DOI: 10.1016/j.ejpe.2014.09.010
    32. 32
      Wardhani, S.; Purwonugroho, D.; Fitri, C. W.; Prananto, Y. P. Effect of PH and Irradiation Time on TiO2-Chitosan Activity for Phenol Photo-Degradation. AIP Conf. Proc. 2018, 2021, 050009,  DOI: 10.1063/1.5062759
    33. 33
      Mestre, A. S.; Carvalho, A. P. Photocatalytic Degradation of Pharmaceuticals Carbamazepine, Diclofenac, and Sulfamethoxazole by Semiconductor and Carbon Materials: A Review. Molecules 2019, 24, 3702,  DOI: 10.3390/molecules24203702
    34. 34
      Beard, M. C.; Luther, J. M.; Nozik, A. J. The Promise and Challenge of Nanostructured Solar Cells. Nat. Nanotechnol. 2014, 9, 951954,  DOI: 10.1038/nnano.2014.292
    35. 35
      Koo, N.; Plachetka, U.; Otto, M.; Bolten, J.; Jeong, J.; Lee, E.; Kurz, H. The Fabrication of a Flexible Mold for High Resolution Soft Ultraviolet Nanoimprint Lithography. Nanotechnology 2008, 19, 225304,  DOI: 10.1088/0957-4484/19/22/225304
    36. 36
      Lemme, M. C.; Moormann, C.; Lerch, H.; Möller, M.; Vratzov, B.; Kurz, H. Triple-Gate Metal–Oxide–Semiconductor Field Effect Transistors Fabricated with Interference Lithography. Nanotechnology 2004, 15, S208S210,  DOI: 10.1088/0957-4484/15/4/016
    37. 37
      AMONIL. High performance UV nanoimprint resist; AMO GmbH. https://www.amo.de/products-services/amonil/ (accessed Dec 03, 2022).
    38. 38
      ISO 10678:2010; International Organization for Standardization. https://www.iso.org/standard/46019.html (accessed May 10, 2023).
    39. 39
      Mills, A. An Overview of the Methylene Blue ISO Test for Assessing the Activities of Photocatalytic Films. Appl. Catal., B 2012, 128, 144149,  DOI: 10.1016/j.apcatb.2012.01.019
  • Supporting Information

    Supporting Information

    ARTICLE SECTIONS
    Jump To

    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsanm.3c02867.

    • Additional experimental details, including: a schematic of the laser interference lithography setup used to fabricate master wafers for nanoimprint lithography; optical absorbance spectra of the nanostructured plasmonic photocatalyst with Au and TiO2 measured on two different panels at three different positions each; and degradation of methylene blue by the nanostructured plasmonic photocatalyst with Au and TiO2 under UV-A with reactor data from measurements on different days (PDF)


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

You’ve supercharged your research process with ACS and Mendeley!

STEP 1:
Click to create an ACS ID

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

MENDELEY PAIRING EXPIRED
Your Mendeley pairing has expired. Please reconnect