Next Article in Journal
Cyclohexene Oxidation with H2O2 over Metal-Organic Framework MIL-125(Ti): The Effect of Protons on Reactivity
Next Article in Special Issue
MoS2/CdS Heterostructure for Enhanced Photoelectrochemical Performance under Visible Light
Previous Article in Journal
Chondroitin Sulfate-Degrading Enzymes as Tools for the Development of New Pharmaceuticals
Previous Article in Special Issue
Enhanced Visible Light Photocatalytic Reduction of Cr(VI) over a Novel Square Nanotube Poly(Triazine Imide)/TiO2 Heterojunction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Characterization and Effect of Ag(0) vs. Ag(I) Species and Their Localized Plasmon Resonance on Photochemically Inactive TiO2

1
Program Studi Ilmu Lingkungan, Universitas Sriwijaya, Jalan Padang Selasa No. 524, Palembang 30121, Sumatera Selatan, Indonesia
2
Leibniz-Institut für Katalyse e. V. (LIKAT), Albert-Einstein-Str. 29a, 18059 Rostock, Germany
3
Arbeitsbereich Medizinische Biologie und Elektronenmikroskopisches Zentrum (EMZ), Universitätsmedizin Rostock, Strempelstr. 14, 18057 Rostock, Germany
4
Laboratory of Inorganic Materials Chemistry, Department of Chemical Engineering and Chemistry, Eindhoven University of Technology, P.O. Box 513, 5600 MB Eindhoven, The Netherlands
5
Program Studi Teknik Kimia, Fakultas Teknik, Universitas Sriwijaya, Jalan Palembang-Inderalaya Km. 32, Inderalaya 30862, Sumatera Selatan, Indonesia
*
Authors to whom correspondence should be addressed.
Catalysts 2019, 9(4), 323; https://doi.org/10.3390/catal9040323
Submission received: 21 February 2019 / Revised: 15 March 2019 / Accepted: 25 March 2019 / Published: 2 April 2019
(This article belongs to the Special Issue Photocatalytic Nanocomposite Materials)

Abstract

:
Commercial TiO2 (anatase) was successfully modified with Ag nanoparticles at different nominal loadings (1%–4%) using a liquid impregnation method. The prepared materials retained the anatase structure and contained a mixture of Ag0 and AgI species. Samples exhibited extended light absorption to the visible region. The effect of Ag loading on TiO2 is studied for the photocatalytic reduction of CO2 to CH4 in a gas–solid process under high-purity conditions. It is remarkable that the reference TiO2 used in this work is entirely inactive in this reaction, but it allows for studying the effect of Ag on the photocatalytic process in more detail. Only in the case of 2% Ag/TiO2 was the formation of CH4 from CO2 observed. Using different light sources, an influence of the localized surface plasmon resonance (LSPR) effect of Ag is verified. A sample in which all Ag has been reduced to the metallic state was less active than the respective sample containing both Ag0 and Ag+, indicating that a mixed oxidation state is beneficial for photocatalytic performance. These results contribute to a better understanding of the effect of metal modification of TiO2 in photocatalytic CO2 reduction.

1. Introduction

In recent decades, atmospheric CO2 concentration has been increasing annually mainly due to the extensive use of fossil fuels [1,2]. Data from NASA show that in April 2018 the concentration of CO2 reached 407.45 ppm, an increase of 29.23% compared to the respective concentration for April 1958 [2]. As a result, the annual mean global surface temperature increased—as CO2 is a major greenhouse gas contributing to climate change—by 0.99 °C, the highest rise observed since 1880 [3]. Many efforts and experimental research have been focused on the capture of CO2 from the atmosphere and/or its conversion to other economically valuable compounds. Since CO2 is a very stable molecule with a C = O direct bond energy of +805 kJ/mol, it requires a high energy input for bond cleavage [4]. One of the most promising technologies to convert CO2 to useful hydrocarbons is the photocatalytic reduction (the chemical transformation of CO2 through the absorption of photons of appropriate wavelength from a semiconducting material—photocatalyst) which offers a solution for both the energy crisis and global warming [5]. The photocatalytic conversion of CO2 to value-added hydrocarbons and renewable fuels, such as methane or methanol, is part of a greater effort in the literature to utilize small molecules such as H2 and CO2, collectively termed as “artificial photosynthesis” [6,7,8,9,10].
As a semiconductor material, titanium dioxide (TiO2) has attracted tremendous interest due to its good photocatalytic activity, low toxicity, chemical stability, abundance in nature, and low cost [11,12]. However, there are several factors which hinder the photocatalytic efficiency of TiO2, such as its absorption range which is limited to wavelengths in the UV region (about 4% of total solar spectrum) [13,14]. Another significant factor reported in many studies is the recombination of charge carriers (electron-hole recombination) in the bulk or the surface of the TiO2 material [15]. Schneider et al. proposed that nearly 90% of all photogenerated electrons recombine rapidly after the electron excitation process. As a result, this phenomenon triggers low quantum yields (<10%) [16].
Therefore, much effort has been focused on enhancing the photocatalytic activity of TiO2, including doping with metal or non-metal species [17,18], surface sensitization [19,20], and noble metal deposition [21,22,23,24,25]. In particular, the noble metal deposition has attracted a lot of attention since it can improve e–h+ separation and enhance the semiconductor’s absorption in the visible light region [26,27]. Noble metals such as Au, Pt, Ag, Pd, Rh, and coinage metals such as Ni etc., have lower Fermi levels than TiO2. Therefore, they can act as electron traps, facilitating the separation of charge carriers and preventing their recombination [26]. On the other hand, the localized surface electron plasmon resonance effect (LSPR) induced by noble metals contributes to the activation of TiO2 absorption in the visible region of the light spectrum.
There are several studies in the literature about photocatalytic CO2 reduction using Ag/TiO2 composites [28,29,30,31,32]. The results vary, depending on the materials’ synthesis and reaction conditions. For example, previous studies revealed that CH3OH is the major CO2 reduction product in liquid-phase experiments, while CH4 is the main product in gas-phase reactions [33]. A gas-phase system offers several advantages like the variation of the concentration of reactants in a wider range without dissolution phenomena [34]. Furthermore, in gas–solid reactor systems, experiments using high purity conditions can be established as a high vacuum can be applied to eliminate gaseous C-containing impurities from the reaction chamber [35]. Performing CO2 reduction experiments using conditions of high-purity is extremely important as any source of carbon-containing species can lead to an overestimation of the products. Due to the high stability of the CO2 molecule, these C-impurities are expected to react faster than CO2, leading to hydrocarbons that do not originate from the actual photocatalytic process facilitated by the tested material.
Furthermore, a photocatalytic CO2 reduction experiment can be performed either in batch or in flow-through mode [35]. On the one hand, flow conditions lead to faster reaction rates. However, since the concentration of even the major photocatalytic CO2 reduction products are quite low, it is very difficult to identify and accurately quantify them during flow-through conditions. On the other hand, a batch mode process allows for a more accurate estimation of the products’ yields even at lower concentrations, as the amount of molecules accumulate over the total reaction time. Besides the use of H2O, both in liquid- and gas-phase, the CO2 photoconversion into CH4 can also be carried out by mediating hydrogen at room temperature and atmospheric pressure called photo-methanation [36]. Furthermore, literature concerning the effect of Ag oxidation state on CO2 photoreduction activities is rarely studied. Most of the research on CO2 photoreduction only focuses on the effect of the Ag metallic co-catalyst which contributes through the electron trapping and LSPR mechanism [33,37].
In this study, Ag nanoparticles were deposited on commercial TiO2 (anatase) using the liquid impregnation method. This method was employed as it is simple, fast, and of low cost [38,39]. The physical and chemical properties of the synthesized Ag/TiO2 composites were extensively studied and characterized using X-ray powder diffraction (XRD), scanning electron microscopy (SEM), ultra violet-visible (UV-Vis) spectrophotometry, X-ray photoelectron spectroscopy (XPS), and transmission electron microscopy (TEM). Furthermore, the assessment of the photocatalytic CO2 reduction activity of the aforementioned materials, the study of the influence of Ag loading, and the effect of mixed Ag species and the LSPR on the products’ concentration was carried out in a high purity gas-phase photoreactor.

2. Results and Discussion

2.1. Synthesis of Photocatalysts

All investigated Ag/TiO2 materials have been successfully prepared by facile impregnation methods. Commercially available TiO2 (anatase, ≥99.0%, Merck) was used in this study, both as a starting and as a reference material. The Ag decorated TiO2 photocatalysts were prepared using a modified liquid impregnation method [14]. The resulting Ag modified TiO2 photocatalysts were referred to as Ag/TiO2X% (X = 1–4), whereby X is the nominal silver loading.

2.2. Characterization of Photocatalysts

2.2.1. Powder X-ray Diffraction

X-ray diffraction techniques were used to determine crystallinity and structural changes which may occur during Ag metal deposition. Figure 1 shows the XRD powder patterns of the pristine TiO2 and the various Ag/TiO2 samples calcined at 500 °C in air atmosphere. As depicted, all prepared materials exhibited high crystallinity. There were no detectable changes in the crystal structure of the materials during Ag deposition. All materials retained the tetragonal crystal structure of TiO2 anatase. The main reflexes at 2θ~25°, 38°, and 48° were attributed to the (101), (004), and (200) planes of TiO2 anatase (JCPDS File No. 83-2243), respectively. For Ag/TiO2 materials, small reflexes at 2θ~44° indicated the presence of Ag0 species with a lattice plane of (200) (JCPDS File No. 04-0783). The emergence of Ag0 species is due to the partial reduction of Ag+ (AgNO3) under high temperature conditions used in the synthetic protocol. The reflex characteristic of the Ag0 species can be clearly seen for the 2%, 3%, and 4% Ag/TiO2 composites. Furthermore, there were also small reflexes (2θ~33.2°) which corresponded to the diffraction pattern of (111) planes of Ag2O species (JCPDS File No. 41-1104). In addition, it can be assumed that Ag was not incorporated into the TiO2 structure, but deposited on its surface (vide infra) [11].
From the powder XRD data it can be concluded that the as-synthesized materials contained a mixture of Ag0 and AgI species, and that the Ag was not completely reduced to Ag0 during the synthesis. This phenomenon occurs due to the calcination in air atmosphere, where O2 inhibits the complete reduction of all Ag species. The calcination process at 500 °C decomposes AgNO3 into different Ag species, NOx gas, and O2 gas.
The average crystallite size of deposited Ag for each sample was calculated from the broadening of the cubic Ag (200) reflex using the Scherrer equation (Equation (1)):
D   =   0.94   λ β 1 / 2 cos θ
where D is the crystallite size, λ is the X-ray wavelength used in this analysis, β1/2 is the full width at half maximum, and θ is the main reflex of XRD pattern. Based on this calculation, the size of deposited Ag nanoparticles varies from 28 nm to 37 nm (Ag/TiO24% exhibited the largest crystallite size among the tested samples). Further SEM and TEM analyses were used to support these data (see Figures 5 and 6).

2.2.2. Electronic Spectra

To study the optical properties and light absorption effects of Ag species deposited on TiO2, UV-Vis spectra were collected from the tested samples in a pressed disk form. Figure 2a demonstrates the absorption properties of the as-synthesized materials in the wavelength range 250–800 nm. As it is clearly seen, the reference TiO2 sample has only UV light absorption. The deposited Ag species contributed to an increase in the material’s absorption towards the visible region as seen in the 420–700 nm wavelength range (see inset graph). This is in accordance with previous work on visible light absorption in Ag nanoparticles species at 410–425 nm due to the surface plasmon resonance effect [40,41,42,43,44,45,46] (Figure 2b). The absorption, with a maximum near 425 nm, can be attributed to the plasmon absorption of spherical Ag nanoparticles [45], whereas the broad absorption reaching down to ~700 nm may be attributed to differently shaped Ag nanoparticles, such as cube-like structures [45], or to the absorption of Ag2O [47].
In order to further investigate the electronic properties of the Ag/TiO2 materials, additional UV-Vis diffuse reflectance spectra were collected for all samples. To this aim, the Kubelka–Munk function, F(R), allows the optical absorbance of the materials to be approximated from their reflectance (R) according to [49] (Figure 3a). The band gap energy of the tested samples can be estimated from the Tauc plot, using equations from the literature [50].
The optical band gap energy calculation was determined by extrapolating the linear region of the plot hv vs (F(R)hv)1/2 as shown in Figure 3b. The calculation shows that the band gap energy of the prepared materials is in the 3.15–3.20 eV range. This slight shift indicates that the deposition of Ag (1%–4%) on the surface of TiO2 does not significantly affect the band gap energy.

2.2.3. Determination and Characterization of Surface Ag on TiO2

X-ray photoelectron spectroscopy was implemented to investigate the surface chemical state and composition of the as-prepared samples. Figure 4a shows the survey spectrum of the 4% Ag/TiO2 sample. From the area Ag 3d and Ti 2p photoelectron peaks, and respective sensitivity factors, the surface Ag:Ti atomic percentage was estimated to be 20%. The positive deviation respect to the nominal Ag:Ti ratio suggests that Ag particles are indeed decorating the TiO2 surface—photoelectron spectroscopy is largely surface sensitive where the Ag is located. Besides the peak from Ti, O, and Ag, traces of C (C 1s) can be detected. This is usually ascribed either to carbon impurities in the titania, or to adventitious carbon during the XPS measurement.
From the high-resolution X-ray photoelectron (XP) spectrum of Ag 3d (Figure 4b), the binding energy (BE) corresponding to the Ag 3d5/2 peak appears at 367.32 eV. In addition, the BE of Ag 3d3/2 is located at 373.32 eV. Those two peaks indicate the formation of AgI species [51]. The peak is asymmetric and broad and therefore it is designated to two Ag species (AgI and metallic Ag). This is in accordance with results from XRD measurements (see above). The asymmetry and broadness disappears in fully reduced samples where metallic Ag is present only. Furthermore, it is noteworthy, that XPS also detected the presence of potassium most probably originating from impurities contained in the precursors used in the synthesis of the materials. The shoulder at high BE in the Ag 3d region is assigned to K 2s. This impurity may contribute to the low photochemical activity because this is known for the comparable impact of sodium in CO2 reduction (see below) [34].
Figure 4c illustrates the Ti 2p spectrum of the same Ag/TiO2 4% sample. The peak for Ti 2p3/2 can be fitted with two components, one located at 458.07 eV and the other at 458.62 eV. Both peaks are attributed to TiIV. The small positive shift (ca. 0.6 eV) is due to the strong interaction between TiO2 and Ag species leading to electron transfer from the former to the latter [52,53]. Another peak located at 463.83 eV is assigned to Ti 2p1/2. The splitting between Ti 2p3/2 and Ti 2p1/2 is 5.75 eV indicates that Ti exists as TiIV species [54]. Meanwhile, no shoulder peaks at BE 462.5 eV and 456.8 eV for Ti 2p1/2 and Ti 2p3/2 core level of TiIII appear, indicating that no TiIII species exist on the surface of the material [55].
Figure 4d depicts the fitting of the O 1s peak. An asymmetric peak of O 1s can be deconvoluted into two peaks. These peaks are located at 529.29 eV and 530.28 eV corresponding to the Ti–O [56] and H–O bonds [57], respectively. All those peaks are corresponding to O2− species.

2.2.4. Electron Microscopy and Energy-Dispersive X-ray Spectroscopy

In order to investigate the surface morphology of the samples, SEM pictures were collected (Figure 5). No significant morphology differences between the pure TiO2 anatase and the Ag/TiO2 can be found (not shown). SEM results show agglomerated spherical TiO2 particles with crystallite diameters in the 100–200 nm range. From Figure 5 it can be clearly seen that the Ag nanoparticles emerge (appearing as white bright spheres) on the surface of TiO2. The size of the Ag nanoparticles varies from 2–50 nm. It is in accordance with calculations from the powder XRD data (see above). The energy dispersive X-ray (EDX) analysis confirmed that the actual Ag loading is 1.1%, 1.8%, 2.1%, and 3.6% for Ag/TiO2 1%, Ag/TiO2 2%, Ag/TiO2 3%, and Ag/TiO2 4%, respectively. Small deviations between nominal and actual loading may be due to the uneven distribution of the Ag nanoparticles on the surface of the Ag-TiO2 composite.
TEM images of the 2% and 4% Ag/TiO2 are shown in Figure 6. As shown in Figure 6a,b, Ag (appearing as small dark spheres) is distributed unevenly on the TiO2 covering only selected areas of the TiO2 surface, which is in accordance with previous research [22]. Furthermore, the size distribution of Ag nanoparticles also varies as seen from Figure 6c. From the TEM images (Figure 6c–e) the lattice fringes allow the identification of crystallographic spacing. The d-spacing value of 0.35 nm corresponds to the anatase (101) plane [57,58], whereas the d values of 0.28 nm and 0.16 nm, respectively, represent the plane of (111) and (220) for Ag2O species [59,60,61]. These results strongly support the results from powder X-ray diffraction (see above).

2.3. Photocatalytic Activity in CO2 Reduction

Sample Testing in Photocatalytic CO2 Reduction

As carbon-containing impurities adsorbed onto the surface of the photocatalysts can lead to overestimation of the concentration of the products of CO2 reduction, special attention must be paid to sufficiently remove them. These impurities might originate from the synthesis process (from precursors, solvents, etc.) or from adsorbed species (e.g., atmospheric CO2 and hydrocarbons). As explained in detail in the experimental section, extensive cleaning of the materials in a gas-phase of only water in inert gas was performed. This cleaning approach usually ensures that impurities are sufficiently removed and that the observed products in photocatalytic CO2 reduction originate predominantly from true photocatalytic reaction between both H2O and CO2. However, this is not the case here. All four Ag loaded samples were tested under conditions of high-purity in photocatalytic CO2 reduction and compared to the reference TiO2 material. Only in the case of Ag/TiO2 2% was a CH4 production not originating from residual C-containing impurities identified. In all other samples (Ag/TiO2 1%, 3%, and 4% and reference TiO2), the levels of CH4 produced in the cleaning procedure (i.e., in absence of CO2) and in the actual reaction with CO2 and H2O in the gas-phase were identical (not shown). So, for all samples except Ag/TiO2 2%, CH4 formation can be attributed solely to left-over impurities. Consequently, only Ag/TiO2 2% is further considered in comparison with the reference TiO2 (Figure 7).
It is remarkable that the reference TiO2 used in this work was almost inactive in photocatalytic CO2 reduction (Figure 7), although a broad-band 200 W Hg/Xe light source was used that was able to excite TiO2. Not only was the CH4 yield very low (~1.2 ppm), it was also identical in the cleaning procedure and the CO2 reduction experiment. The low to non-existent activity may have to do with TiO2 being present only in anatase structure, because usually P25 with both rutile and anatase components (~80/20%) is used and shows high activity [32,34,35,62]. Furthermore, the titania may be highly defective, so that recombination events are predominant. While we cannot clearly give an answer to the inactivity of the reference TiO2, it is obvious that the presence of Ag induces photocatalytic activity, the origin of which is further studied.
The 2% Ag loaded sample exhibited considerably higher CH4 production compared to the bare anatase reference TiO2 (Figure 7). Furthermore, almost twice as much CH4 was formed, as in the case of the cleaning procedure, so at least half of all products must originate from CO2 as carbon source. The higher photocatalytic efficiency of the Ag decorated TiO2 sample is often attributed to the fact that the Ag species act as electron scavengers and therefore the recombination of the photo-generated e–h+ pairs is hindered [63]. The photo-generated electron transfer from the TiO2 conduction band to the Ag species deposited on the surface is thermodynamically possible due to the higher Fermi energy level of TiO2 compared to Ag [64]. As a result, the formation of a Schottky barrier occurs at the interface between Ag and TiO2, which can further improve the charge separation and consequently, the photocatalytic activity [65]. The Schottky barrier prevents the back-transfer of electrons from Ag to TiO2, thus they remain trapped in the Ag species. Since the titania itself is inactive in the present work, two explanations may account for the improved activity of the Ag/TiO2 sample: Either recombination in the bare titania is so much favored that charge separation is only possible in presence of Ag, or titania contributes nothing to the observed photoactivity, because it is only due to the localized surface plasmon resonance (LSPR) effect. The LSPR effect induced by the presence of Ag nanoparticles in the visible region can boost the energy of trapped electrons through the strong local electron field of LSPR, making it react with electron acceptors more easily [28]. This synergistic effect of the LSPR effect and the e–h+ separation can lead to an enhancement in the photocatalytic activity as shown in Figure 2b.
In order to test how the LSPR effect influences the CH4 production, a CO2 reduction experiment was performed with the Ag/TiO2 2% sample, using a 10 W 365 nm high-power LED while all the other experimental parameters remained the same. As it can be seen from Figure 7, the CH4 concentration is much higher when the Hg/Xe lamp (containing wavelengths in both UV and visible range of the light spectrum) was used compared to the CH4 production under LED illumination. This means that the absorption of the Ag/TiO2 2% sample in the 420–550 nm range due to the LSPR effect is the main reason for the observed photocatalytic activity of this sample.
The observation of favorable performance of a sample with ~2 wt % Ag is in a good agreement with Zhao et al. [37]. They reported that the optimal Ag concentration on TiO2 was 2 wt % in simultaneous H2 production and photocatalytic CO2 reduction. Similar results were also obtained by Liu et al. [28] and Yu et al. [30]. Based on Liu et al. the composite of 2.5% Ag/TiO2 exhibits the best activity in CO2 reduction compared with other Ag loadings. On the other hand, Yu et al. proposed that 1.5% Ag/TiO2 is the best material. This is due to the further increase of Ag loading triggering the LSPR peak to become slightly red-shifted and broader, implying that the Ag nanoparticles begin to agglomerate. However, the exact CH4 production rate for those studies was different due to the difference of both material properties and reactor set up.
Additionally, it must be clarified which oxidation state of Ag is required for high photoactivity. In order to study the effect of the Ag oxidation state, a sample in which all Ag has been reduced to the metallic state (namely Ag0/TiO2 2%) was tested in CO2 reduction under similar conditions as the Ag/TiO2 2%, which contained both Ag0 and AgI. As it can be seen from Figure 7, a mixed oxidation state is photocatalytically beneficial. One explanation for this higher photocatalytic activity may be the presence of a Ag2O–TiO2 interface that may act as an e- trapping layer compared to the pure Ag0 loaded sample. As reported in [66] an interfacial charge transfer (IFCT) takes place between Ag2O and TiO2 under band gap irradiation. This is facilitated by the position of the energy bands of Ag2O and TiO2, as the conduction band of Ag2O is more negative than that of TiO2. This electron transfer hinders the e–h+ recombination in the Ag2O. The photo-produced h+ reacts with H2O generating OH° and H+.
Furthermore, considering the function of the Ag0/AgI species as a co-catalyst, mixed systems composed of a metallic core and an oxide shell have often been found beneficial for different reactions. For example, a Ni/NiO core shell system, as well as Rh/Cr2O3 or CuOx/Cr2O3 systems (in which the core-shell structure is debated) have been found beneficial for water splitting, particularly for hydrogen evolution [67,68,69,70]. While the metal ensures a conducting contact with the semiconductor and may act as the catalytic active site for reduction [70], the oxide (shell) may have other functions, such as prevention of the back reaction or functioning for water oxidation [67,71]. Although the structural arrangement of the Ag0 and AgI (core-shell, neighboring, far from each other, etc.) is not known in the present case, it is possible that both species adopt important roles in the mechanism. We can only speculate that electron conduction and electron pooling may proceed on Ag0, whereas adsorption of CO2 or stabilization of some relevant intermediates may take place on AgI. This needs to be clarified in future studies.
Since a general mechanism for photocatalytic CO2 reduction is not yet clear, an a priori design of a suitable catalyst is not an easy task. However, TiO2-based materials with co-catalysts (mixed states, metal loaded, z-schemes, etc.) have the potential to exhibit good efficiencies in producing useful chemicals from CO2. The optimum synthesis parameters should be optimized both in terms of efficiency, but also in terms of cost, environmental burden, and sustainability. The necessity for high purity experimental conditions is eminent in combination with (at least) blank experiments in order to reach safer conclusions about novel materials.

3. Experimental Section

3.1. General Considerations

Powder XRD data were obtained on a Miniflex 600 device (Rigaku, Tokyo, Japan) in the range of 2θ = 3–80° (0.2° step, 30 kV, 10 mA, Cu Kα). UV-Vis spectra were recorded on a UV-2450PC PharmaSpec device (Shimadzu, Kyoto, Japan) for dry-pressed sample disks and on a Lambda 365 device (PerkinElmer, Waltham, MA, USA) in the diffuse reflectance mode for the as-synthesized powders, respectively. XPS data were recorded on a K-Alpha device (ThermoFisher Scientific, Waltham, MA, USA) using a 72 W monochromated Al Kα source. All XP spectra were referenced according to the C 1 s peak at 285 eV. Simulations of the experimental photopeaks were carried out using a mixed Gaussian/Lorentzian peak fit procedure according to the CasaXPS software. Semi-quantitative analysis accounted for a nonlinear Shirley background subtraction. SEM measurements were done using a Merlin VP compact device (Zeiss, Oberkochen, Germany) and additional EDX investigations were performed using a Quantax 400 device (Bruker, Billerica, MA, USA). TEM images were recorded either on a JEM-1400 device (Jeol, Akishima, Japan), or on a JEM-ARM200F device (Jeol, Akishima, Japan). Gas chromatograph (GC) measurements during photocatalytic CO2 reduction were collected using a Tracera GC-2010 Plus device (Shimadzu, Berlin, Germany) equipped with a barrier ionization discharge detector (BID) and a flame ionization detector (FID).

3.2. Materials

All starting materials were purchased commercially from Merck and used as received: Titania TiO2 (anatase, ≥99.0%), silver nitrate (AgNO3, ≥99.0%).

3.3. Preparation of Photocatalysts

The Ag decorated TiO2 photocatalysts were prepared using a liquid impregnation method, which is exemplified in the following for the 1 mol% Ag/TiO2 sample. TiO2 powder (7.99 g) and 0.17 g of AgNO3 precursor were added in 100 mL of distilled water. The theoretical loading amount of Ag was 1 mol% (the molar ratio of Ag to TiO2). The slurry was kept under stirring and heated at 80 °C for 4 h to achieve Ag+ equilibrium on the surface of TiO2. The stirring process was equipped with reflux apparatus. Then, the slurry was allowed to settle at room temperature overnight. The formed slurry was dried in an oven for 12 h at a temperature of 100 °C. The resulting solid was ground with a mortar and subsequently calcined at 500 °C for 3 h in air. For the preparation of the samples with different Ag loadings, the amount of AgNO3 was varied accordingly, while all other reagent amounts were kept similar. The resulting Ag modified TiO2 photocatalysts are referred to as Ag/TiO2X% (X = 1–4) [30]. The Ag0/TiO2 2% composite was prepared by reducing the Ag/TiO2 2% sample using 20% H2 (in Ar) for 1 h at 200 °C.

3.4. Photocatalytic Setup

The photocatalytic CO2 reduction experiments were carried out in a high-purity gas-phase photoreactor setup. A detailed description of the photocatalytic reactor design was given by Mei et al. [6]. In brief, the photoreactor is made exclusively of stainless steel components suitable for ultra-high vacuum applications. All of the tube connections were realized using vacuum coupling radiation (VCR) fittings, thus ensuring a grease-free sealing of all interconnecting parts. The sample powder was spread as evenly as possible on a quartz sample holder (diameter ~2.5 cm) placed in the center of the photoreactor. High purity He (6.0) and 1.5% CO2/He (5.5/6.0) mixture were used in the experiments and their respective flows towards the reactor were adjusted using a set of two mass flow controllers (MFC). Water enrichment of the gas mixtures was performed in a temperature-controlled stainless-steel saturator. All water-carrying pipes were constantly heated to 120 °C to avoid internal H2O condensation. A 200 W Hg/Xe lamp (Newport Oriel) emitting UV and visible light was used as the irradiation source, as well as a 10 W array of 4 high-power LEDs emitting at 365 nm (SeoulVioSys). A water-based filter was used for IR radiation removal from the lamp to avoid sample heating. The product analysis was performed using a Shimadzu Tracera GC 2010 plus gas chromatograph (GC) equipped with a barrier ionization discharge detector (BID—Q Bond 50 m column 5A molecular sieve) and a flame ionization detector (FID—Alumina Bond/KCl 50 m column). Helium (He) at a flow of 25 mL min−1 was used as the carrier gas in the GC. The temperature of the columns was increased from initially 30 °C up to finally 160 °C (heating rate: 20 °C min−1). Then the temperature was decreased back to 30 °C (cooling rate: 50 °C min−1) and immediately afterwards increased again to 160 °C (heating rate: 20 °C min−1). In total, 33 min were needed for a complete measurement with a 5-min cool-down period. This procedure facilitated desorption of molecules from the column.

3.5. Sample Pretreatment

Before studying the performance of the samples in photocatalytic CO2 reduction, extensive cleaning was performed to clean the materials’ surface from residual carbon-containing species. These impurities can react with CO2 and H2O molecules to form hydrocarbons, thus leading to overestimation of the efficiency of the materials towards CO2 reduction. The cleaning process applied involved two steps: In the first step, the samples were re-calcined at 400 °C for 3 h at a rate of 180 K h−1 under a synthetic air environment to remove the majority of the adsorbed C-containing impurities. In the second step, the cleaning process continued with the sample (50 mg) inside the photoreactor in batch mode. The reactor was filled with H2O (TH2O = 5 °C)-saturated pure He (at a flow of 10 mL min−1) up to a final pressure of 1500 mbar. A gas sample was collected every 45 min from the batch reactor, for a total duration of 6 h, with the first chromatogram taken without light irradiation. As with each collected sample, when the pressure inside the reactor dropped, a pressure-based correction was performed. Subsequently, a flow-through cleaning process took place, where moist pure He flowed continuously (at a rate of 20 mL min−1) over the sample under light irradiation for 3 h. Under these circumstances, detection of products from sample cleaning was not accurately possible, so the flow cleaning was followed by a second batch cleaning process. The amount of CH4 produced from impurities in this second batch cleaning were compared to CH4 formation in the photocatalytic CO2 reduction experiment in order to evaluate the relative number of products truly originating from CO2.

3.6. Photocatalytic CO2 Reduction

The photocatalytic CO2 reduction experiment was carried out in batch mode conditions. The experimental conditions and GC analysis were similar to the batch cleaning process, except for the presence of CO2 as a reactant gas. Thus, the gas-phase feed of the reactor consisted of 1.5% CO2 diluted in He and saturated with H2O at 5 °C. The product analysis was also performed every 45 min with the necessary pressure drop corrections. An extensive set of blank measurements (no photocatalyst, no CO2, no light, and combinations of the above) were performed to ensure that no, or sufficiently low, concentration of hydrocarbons were produced at conditions other than a CO2 reduction experiment.

4. Conclusions

In this study, Ag nanoparticles deposited on commercial anatase TiO2 were successfully prepared by liquid impregnation method. As XPS analysis results indicated, the composite contained the mixture of AgI and Ag0 on the surface of TiO2. It was found that all silver/titania materials had visible light absorption due to surface plasmon absorption of electrons in silver species. The TEM and SEM images showed the uneven dispersion of Ag nanoparticles with a diameter of ~2–50 nm on the TiO2 surface. Only the sample with 2 wt % Ag was revealed to be an effective material for CO2 reduction. The LSPR effect was studied using two different irradiation sources containing only UV (365 nm LED) or both UV and visible (Hg/Xe lamp) wavelengths, and it was identified as the main reason for the photocatalytic activity of the Ag/TiO2 2% sample. In addition, the presence of mixed-valent silver was beneficial for CH4 formation in photocatalytic CO2 reduction. A sample in which all silver has been reduced to the metallic state was instead less active.

Author Contributions

Conceptualization, C.T.H., N.G.M., and T.P.; methodology, C.T.H., N.G.M., T.P., F.E.O., and A.S.; validation, N.G.M., T.P., J.S., M.D.B., B.Y., and F.G.; formal analysis, C.T.H., N.G.M., T.P., F.E.O., A.S., M.D.B, and B.Y.; investigation, C.T.H., N.G.M., T.P., F.E.O., A.S., and A.H.; writing—original draft preparation, C.T.H., N.G.M., T.P., and A.H.; writing—review and editing, N.G.M., T.P., J.S., and F.G.; visualization, C.T.H., N.G.M., T.P., F.E.O., and A.S.; supervision, J.S., and F.G.; project administration, J.S., T.P., and F.G.

Funding

This research was funded by the Ministry of Research, Technology, and Higher Education of Indonesia, grant number: 2064.31/D3/PG2017. Part of this research was funded by the German Ministry of Education and Research (BMBF) in the scope of the funding scheme CO2Plus, Project-No. 033RC033, PROPHECY.

Acknowledgments

The authors thank M.-M. Pohl and C. Kreyenschulte (LIKAT, Rostock) for obtaining TEM images, and A. Simmula and A. Lehmann (LIKAT, Rostock) for elemental analysis.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Low, J.; Cheng, B.; Yu, J. Surface modification and enhanced photocatalytic CO2 reduction performance of TiO2: A review. Appl. Surf. Sci. 2017, 392, 658–686. [Google Scholar] [CrossRef]
  2. NASA. Climate Change Vital Signs of the Planet: Carbon Dioxide. Available online: https://climate.nasa.gov/vital-signs/carbon-dioxide/ (accessed on 6 March 2018).
  3. NASA. Global Climate Change Vital Signs of the Planet: Global Temperature. Available online: https://climate.nasa.gov/vital-signs/global-temperature/ (accessed on 6 March 2018).
  4. Jiang, Z.; Xiao, T.; Kuznetsov, V.L.; Edwards, P.P. Turning carbon dioxide into fuel. Philos. Trans. A Math. Phys. Eng. Sci. 2010, 368, 3343–3364. [Google Scholar] [CrossRef] [Green Version]
  5. Ji, Y.; Luo, Y. New mechanism for photocatalytic reduction of CO2 on the anatase TiO2 (101) surface: The essential role of oxygen vacancy. J. Am. Chem. Soc. 2016, 138, 15896–15902. [Google Scholar] [CrossRef]
  6. Mei, B.; Pougin, A.; Strunk, J. Influence of photodeposited gold nanoparticles on the photocatalytic activity of titanate species in the reduction of CO2 to hydrocarbons. J. Catal. 2013, 306, 184–189. [Google Scholar] [CrossRef]
  7. Li, K.; An, X.; Park, K.H.; Khraisheh, M.; Tang, J. A critical review of CO2 photoconversion: Catalysts and reactors. Catal. Today 2014, 224, 3–12. [Google Scholar] [CrossRef]
  8. Handoko, A.D.; Li, K.; Tang, J. Recent progress in artificial photosynthesis: CO2 photoreduction to valuable chemicals in a heterogeneous system. Curr. Opin. Chem. Eng. 2012, 2, 200–206. [Google Scholar] [CrossRef]
  9. Yuan, L.; Xu, Y.-J. Photocatalytic conversion of CO2 into value-added and renewable fuels. Appl. Surf. Sci. 2015, 342, 154–167. [Google Scholar] [CrossRef]
  10. Sato, S.; Arai, T.; Morikawa, T. Toward solar-driven photocatalytic CO2 reduction using water as an electron donor. Inorg. Chem. 2015, 54, 5105–5113. [Google Scholar] [CrossRef]
  11. Zhao, B.; Chen, Y.-W. Ag/TiO2 sol prepared by a sol–gel method and its photocatalytic activity. J. Phys. Chem. Solids 2011, 72, 1312–1318. [Google Scholar] [CrossRef]
  12. Lai, Y.; Chen, Y.; Zhuang, H.; Lin, C. A facile method for synthesis of Ag/TiO2 nanostructures. Mater. Lett. 2008, 62, 3688–3690. [Google Scholar] [CrossRef]
  13. Li, X.; Wen, J.; Low, J.; Fang, Y.; Yu, J. Design and fabrication of semiconductor photocatalyst for photocatalytic reduction of CO2 to solar fuel. Sci. China Mater. 2014, 57, 70–100. [Google Scholar] [CrossRef]
  14. Saha, S.; Wang, J.M.; Pal, A. Nano silver impregnation on commercial TiO2 and a comparative photocatalytic account to degrade malachite green. Sep. Purif. Technol. 2012, 89, 147–159. [Google Scholar] [CrossRef]
  15. Etacheri, V.; Valentin, C.D.; Schneider, J.; Bahnemann, D.; Pillai, S.C. Visible-light activation of TiO2 photocatalysts: Advances in theory and experiments. J. Photochem. Photobiol. C 2015, 25, 1–83. [Google Scholar] [CrossRef]
  16. Schneider, J.; Matsuoka, M.; Takeuchi, M.; Zhang, J.; Horiuchi, Y.; Anpo, M.; Bahnemann, D.W. Understanding TiO2 Photocatalysis: Mechanisms and Materials. Chem. Rev. 2014, 114, 9919–9986. [Google Scholar] [CrossRef] [PubMed]
  17. Ablat, A.; Wu, R.; Mamat, M.; Ghupur, Y.; Aimidula, A.; Bake, M.A.; Gholam, T.; Wang, J.; Qian, H.; Wu, R.; et al. Electronic structure and room temperature ferromagnetism of C doped TiO2. Solid State Commun. 2016, 243, 7–11. [Google Scholar] [CrossRef]
  18. Kočí, K.; Matějová, L.; Ambrožová, N.; Šihor, M.; Troppová, I.; Čapek, L.; Kotarba, A.; Kustrowski, P.; Hospodková, A.; Obalová, L. Optimization of cerium doping of TiO2 for photocatalytic reduction of CO2 and photocatalytic decomposition of N2O. J. Sol-Gel Sci. Technol. 2016, 78, 550–558. [Google Scholar] [CrossRef]
  19. Rho, W.-Y.; Jeon, H.; Kim, H.-S.; Chung, W.-J.; Suh, J.S.; Jun, B.-H. Recent progress in dye-sensitized solar cells for improving efficiency: TiO2 nanotube arrays in active layer. J. Nanomater. 2015, 2015, 1–17. [Google Scholar] [CrossRef]
  20. Wang, T.; Xia, M.; Kong, X. The pros and cons of polydopamine-sensitized titanium oxide for the photoreduction of CO2. Catalysts 2018, 8, 215. [Google Scholar] [CrossRef]
  21. Yoon, J.-W.; Sasaki, T.; Koshizaki, N.; Traversa, E. Preparation and characterization of M/TiO2 (M = Ag, Au, Pt) nanocomposite thin films. Scripta Mater. 2001, 44, 1865–1868. [Google Scholar] [CrossRef]
  22. Kim, K.D.; Han, D.N.; Lee, J.B.; Kim, H.T. Formation and characterization of Ag-deposited TiO2 nanoparticles by chemical reduction method. Scripta Mater. 2006, 54, 143–146. [Google Scholar] [CrossRef]
  23. Tahir, K.; Ahmad, A.; Li, B.; Khan, A.U.; Nazir, S.; Khan, S.; Khan, Z.U.H.; Khan, S.U. Preparation, characterization and an efficient photocatalytic activity of Au/TiO2 nanocomposite prepared by green deposition method. Mater. Lett. 2016, 178, 56–59. [Google Scholar] [CrossRef]
  24. Dozzi, M.V.; Brocato, S.; Marra, G.; Tozzola, G.; Meda, L.; Selli, E. Aqueous ammonia abatement on Pt- and Ru-modified TiO2: Selectivity effects of the metal nanoparticles deposition method. Catal. Today 2017, 287, 148–154. [Google Scholar] [CrossRef]
  25. Olivo, A.; Ghedini, E.; Pascalicchio, P.; Manzoli, M.; Cruciani, G.; Signoretto, M. Sustainable carbon dioxide photoreduction by a cooperative effect of reactor design and titania metal promotion. Catalysts 2018, 8, 41. [Google Scholar] [CrossRef]
  26. Mital, G.S.; Manoj, T. A review of TiO2 nanoparticles. Chin. Sci. Bull. 2011, 56, 1639–1657. [Google Scholar] [CrossRef]
  27. Eskandarloo, H.; Badiei, A.; Behnajady, M.A.; Afshar, M. Enhanced photocatalytic removal of phenazopyridine by using silver-impregnated SiO2–TiO2 nanoparticles: Optimization of synthesis variables. Res. Chem. Intermediat. 2015, 41, 9929–9949. [Google Scholar] [CrossRef]
  28. Liu, E.; Kang, L.; Wu, F.; Sun, T.; Hu, X.; Yang, Y.; Liu, H.; Fan, J. Photocatalytic reduction of CO2 into methanol over Ag/TiO2 nanocomposites enhanced by surface plasmon resonance. Plasmonics 2014, 9, 61–70. [Google Scholar] [CrossRef]
  29. Kočí, K.; Zatloukalová, K.; Obalová, L.; Krejčíková, S.; Lacný, Z.; Čapek, L.; Hospodková, A.; Šolcová, O. Wavelength effect on photocatalytic reduction of CO2 by Ag/TiO2 catalyst. Chin. J. Catal. 2011, 32, 812–815. [Google Scholar] [CrossRef]
  30. Kulkarni, R.M.; Malladi, R.S.; Hanagadakar, M.S.; Doddamani, M.R.; Bhat, U.K. Ag-TiO2 nanoparticles for photocatalytic degradation of lomefloxacin. Desalin. Water Treat. 2016, 57, 1–8. [Google Scholar] [CrossRef]
  31. Kočí, K.; Matějů, K.; Obalová, L.; Krejčíková, S.; Lacný, Z.; Plachá, D.; Čapek, L.; Hospodková, A.; Šolcová, O. Effect of silver doping on the TiO2 for photocatalytic reduction of CO2. Appl. Catal. B-Environ. 2010, 96, 239–244. [Google Scholar] [CrossRef]
  32. Dilla, M.; Pougin, A.; Strunk, J. Evaluation of the plasmonic effect of Au and Ag on Ti-based photocatalysts in the reduction of CO2 to CH4. J. Energy Chem. 2017, 26, 277–283. [Google Scholar] [CrossRef]
  33. Yu, B.; Zhou, Y.; Li, P.; Tu, W.; Li, P.; Tang, L.; Ye, J.; Zou, Z. Photocatalytic reduction of CO2 over Ag/TiO2 nanocomposites prepared with a simple and rapid silver mirror method. Nanoscale 2016, 8, 11870–11874. [Google Scholar] [CrossRef]
  34. Pougin, A.; Dilla, M.; Strunk, J. Identification and exclusion of intermediates of photocatalytic CO2 reduction on TiO2 under conditions of highest purity. Phys. Chem. Chem. Phys. 2016, 18, 10809–10817. [Google Scholar] [CrossRef]
  35. Dilla, M.; Schlögl, R.; Strunk, J. Photocatalytic CO2 reduction under continuous flow high-purity conditions: Quantitative evaluation of CH4 formation in the steady state. ChemCatChem 2017, 9, 696–704. [Google Scholar] [CrossRef]
  36. Tampi, K.R.; Kiwi, J.; Grätzel, M. Methanation and photo-methanation of carbon dioxide at room temperature and atmospheric pressure. Nature 1987, 327, 506–508. [Google Scholar] [CrossRef]
  37. Zhao, C.; Krall, A.; Zhao, H.; Zhang, Q.; Li, Y. Ultrasonic spray pyrolysis synthesis of Ag/TiO2 nanocomposite photocatalysts for simultaneous H2 production and CO2 reduction. Int. J. Hydrog. Energy 2012, 37, 9967–9976. [Google Scholar] [CrossRef]
  38. Thein, M.T.; Chim, J.E.; Punga, S.-Y.; Pung, Y.-F. Highly UV light driven WOx@ZnO nanocomposites synthesized by liquid impregnation method. J. Ind. Eng. Chem. 2017, 46, 119–129. [Google Scholar] [CrossRef]
  39. Amadio, T.d.M.; Hotza, D.; Neto, J.B.R.; Blosi, M.; Costa, A.L.; Dondi, M. Bentonites functionalized by impregnation with TiO2, Ag, Pd and Au nanoparticles. Appl. Clay Sci. 2017, 146, 1–6. [Google Scholar] [CrossRef]
  40. Lee, K.-C.; Lin, S.-J.; Lin, C.-H.; Lin, C.-H.; Tsai, C.-S.; Lu, Y.-J. Size effect of Ag nanoparticles on surface plasmon resonance. Surf. Coat. Tech. 2008, 202, 5339–5342. [Google Scholar] [CrossRef]
  41. Leong, K.H.; Gan, B.L.; Ibrahim, S.; Saravanan, P. Synthesis of surface plasmon resonance (SPR) triggered Ag/TiO2 photocatalyst for degradation of endocrine disturbing compounds. Appl. Surf. Sci. 2014, 319, 128–135. [Google Scholar] [CrossRef]
  42. Handoko, C.T.; Huda, A.; Bustan, M.D.; Yudono, B.; Gulo, F. Green synthesis of silver nanoparticle and its antibacterial activity. Rasayan J. Chem. 2017, 10, 1137–1144. [Google Scholar] [CrossRef]
  43. Eom, H.; Jung, J.-Y.; Shin, Y.; Kim, S.; Choi, J.-H.; Lee, E.; Jeong, J.-H.; Park, I. Strong localized surface plasmon resonance effects of Ag/TiO2 core–shell nanowire arrays in UV and visible light for photocatalytic activity. Nanoscale 2014, 6, 226–234. [Google Scholar] [CrossRef]
  44. Handoko, C.T.; Huda, A.; Gulo, F. Synthesis pathway and powerful antimicrobial properties of silver nanoparticle: A critical review. Asian J. Sci. Res. 2019, 12, 1–17. [Google Scholar] [CrossRef]
  45. Linic, S.; Christopher, P.; Ingram, D.B. Plasmonic-metal nanostructures for efficient conversion of solar to chemical energy. Nat. Mater. 2011, 10, 911–921. [Google Scholar] [CrossRef]
  46. Boerigter, C.; Aslam, U.; Linic, S. Mechanism of charge transfer from plasmonic nanostructures to chemically attached materials. ACS Nano 2016, 10, 6108–6115. [Google Scholar] [CrossRef]
  47. Shi, L.; Liang, L.; Ma, J.; Wang, F.; Sun, J. Enhanced photocatalytic activity over the Ag2O–g-C3N4 composite under visible light. Catal. Sci. Technol. 2014, 4, 758–765. [Google Scholar] [CrossRef]
  48. Wu, F.; Hu, X.; Fan, J.; Liu, E.; Sun, T.; Kang, L.; Hou, W.; Zhu, C.; Liu, H. Photocatalytic activity of Ag/TiO2 nanotube arrays enhanced by surface plasmon resonance and application in hydrogen evolution by water splitting. Plasmonics 2012, 8, 501–508. [Google Scholar] [CrossRef]
  49. Jahan, F.; Islam, M.H.; Smith, B.E. Band gap and refractive index determination of Mo-black coatings using several techniques. Sol. Energy Mater. Sol. Cells 1995, 37, 283–293. [Google Scholar] [CrossRef]
  50. Uddin, M.T.; Nicolas, Y.; Olivier, C.; Jaegermann, W.; Rockstroh, N.; Junge, H.; Toupance, T. Band alignment investigations of heterostructure NiO/TiO2 nanomaterials used as efficient heterojunction earth-abundant metal oxide photocatalysts for hydrogen production. Phys. Chem. Chem. Phys. 2017, 19, 19279–19288. [Google Scholar] [CrossRef] [PubMed]
  51. Belver, C.; Hinojosa, M.; Bedia, J.; Tobajas, M.; Alvarez, M.A.; Rodríguez-González, V.; Rodriguez, J.J. Ag-coated heterostructures of ZnO-TiO2/delaminated montmorillonite as solar photocatalysts. Materials 2017, 10, 960. [Google Scholar] [CrossRef] [PubMed]
  52. Wang, S.; Qian, H.; Hu, Y.; Dai, W.; Zhong, Y.; Chen, J.; Hu, X. Facile one-pot synthesis of uniform TiO2–Ag hybrid hollow spheres with enhanced photocatalytic activity. Dalton Trans. 2013, 42, 1122–1128. [Google Scholar] [CrossRef]
  53. Duan, Y.; Zhang, M.; Wang, L.; Wang, F.; Yang, L.; Li, X.; Wang, C. Plasmonic Ag-TiO2−x nanocomposites for the photocatalytic removal of NO under visible light with high selectivity: The role of oxygen vacancies. Appl. Catal. B-Environ. 2017, 204, 67–77. [Google Scholar] [CrossRef]
  54. Lei, T.; Xie, Y.; Wang, X.; Miao, S.; Xiong, J.; Yan, C. TiO2 feather duster as effective polysulfides restrictor for enhanced electrochemical kinetics in lithium–sulfur batteries. Small 2017, 13, 1–9. [Google Scholar] [CrossRef] [PubMed]
  55. Wang, Z.; Yang, C.; Lin, T.; Yin, H.; Chen, P.; Wan, D.; Xu, F.; Huang, F.; Lin, J.; Xie, X.; et al. Visible-light photocatalytic, solar thermal and photoelectrochemical properties of aluminium-reduced black titania. Energy Environ. Sci. 2017, 6, 3007–3014. [Google Scholar] [CrossRef]
  56. Wang, Z.; Yang, C.; Lin, T.; Yin, H.; Chen, P.; Wan, D.; Xu, F.; Huang, F.; Lin, J.; Xie, X.; et al. H-doped black titania with very high solar absorption and excellent photocatalysis enhanced by localized surface plasmon resonance. Adv. Funct. Mater. 2013, 23, 5444–5450. [Google Scholar] [CrossRef]
  57. Wang, F.; Zhang, S.; Li, C.; Liu, J.; He, S.; Zhao, Y.; Yan, H.; Wei, M.; Evans, D.G.; Duan, X. Catalytic behavior of supported Ru nanoparticles on the (101) and (001) facets of anatase TiO2. RSC Adv. 2014, 4, 10834–10840. [Google Scholar] [CrossRef]
  58. Chowdhury, I.H.; Ghosh, S.; Naskar, M.K. Aqueous-based synthesis of mesoporous TiO2 and Ag–TiO2 nanopowders for efficient photodegradation of methylene blue. Ceram. Int. 2016, 42, 2488–2496. [Google Scholar] [CrossRef]
  59. Chen, L.; Hua, H.; Yang, Q.; Hu, C. Visible-light photocatalytic activity of Ag2O coated Bi2WO6 hierarchical microspheres assembled by nanosheets. Appl. Surf. Sci. 2015, 327, 62–67. [Google Scholar] [CrossRef]
  60. Rokade, A.A.; Patil, M.P.; Yoo, S.I.; Lee, W.K.; Park, S.S. Pure green chemical approach for synthesis of Ag2O nanoparticles. Green Chem. Lett. Rev. 2016, 9, 216–222. [Google Scholar] [CrossRef]
  61. Li, M.Y.; Mao, Y.Q.; Yang, S.K.; Dai, T.T.; Yang, H.; Feng, F.; Wu, T.; Chen, M.; Xu, G.Q.; Wu, J.H. Out-of-Substrate Ag−Ag2O nanoplates: Surfactantless photochemical synthesis, structural evolution, and mechanistic study. ACS Omega 2016, 1, 696–705. [Google Scholar] [CrossRef]
  62. Dilla, M.; Mateblowski, A.; Ristig, S.; Strunk, J. Photocatalytic CO2 reduction under continuous flow high-purity conditions: Influence of light intensity and H2O concentration. ChemCatChem 2017, 9, 4345–4352. [Google Scholar] [CrossRef]
  63. Nasirian, M.; Mehrvar, M. Modification of TiO2 to enhance photocatalytic degradation of organics in aqueous solutions. J. Environ. Chem. Eng. 2016, 4, 4072–4082. [Google Scholar] [CrossRef]
  64. Sclafani, A.; Herrmann, J.-M. Influence of metallic silver and of platinum-silver bimetallic deposits on the photocatalytic activity of titania (anatase and rutile) in organic and aqueous media. J. Photochem. Photobiol. A 1998, 113, 181–188. [Google Scholar] [CrossRef]
  65. Linsebigler, A.L.; Lu, G.; John, T.; Yates, J. Photocatalysis on TiO2 surfaces: Principles, mechanisms, and selected results. Chem. Rev. 1995, 95, 735–758. [Google Scholar] [CrossRef]
  66. Rtimi, S.; Baghriche, O.; Sanjines, R.; Pulgarin, C.; Bensimon, M.; Kiwi, J. TiON and TiON-Ag sputtered surfaces leading to bacterial inactivation under indoor actinic light. J. Photochem. Photobiol. A 2013, 256, 52–63. [Google Scholar] [CrossRef] [Green Version]
  67. Busser, G.W.; Schuhmann, W.; Mei, B.; Pougin, A.; Willinger, M.-G.; Strunk, J.; Schlögl, R.; Gutkowski, R.; Muhler, M. Photodeposition of copper and chromia on gallium oxide: The role of co-catalysts in photocatalytic water splitting. ChemSusChem 2014, 7, 1030–1034. [Google Scholar] [CrossRef] [PubMed]
  68. Phivilay, S.P.; Roberts, C.A.; Puretzky, A.A.; Domen, K.; Wachs, I.E. Fundamental bulk/surface structure−photoactivity relationships of supported (Rh2−yCryO3)/GaN photocatalysts. J. Phys. Chem. Lett. 2013, 4, 3719–3724. [Google Scholar] [CrossRef]
  69. Maeda, K.; Teramura, K.; Lu, D.; Saito, N.; Inoue, Y.; Domen, K. Roles of Rh/Cr2O3 (core/shell) nanoparticles photodeposited on visible-light-responsive (Ga1−xZnx)(N1−xOx) solid solutions in photocatalytic overall water splitting. J. Phys. Chem. C 2007, 111, 7554–7560. [Google Scholar] [CrossRef]
  70. Domen, K.; Kudo, A.; Onishi, T. Mechanism of photocatalytic decomposition of water into H2 and O2 over NiO-SrTiO3. J. Catal. 1986, 102, 92–98. [Google Scholar] [CrossRef]
  71. Yoshida, M.; Takanabe, K.; Maeda, K.; Ishikawa, A.; Kubota, J.; Sakata, Y.; Ikezawa, Y.; Domen, K. Role and function of noble-metal/Cr-layer core/shell structure cocatalysts for photocatalytic overall water splitting studied by model electrodes. J. Phys. Chem. C 2009, 113, 10151–10157. [Google Scholar] [CrossRef]
Figure 1. Powder XRD patterns of pristine TiO2 and Ag/TiO2 materials. Small squares (□) at 2θ = 33.2° indicate the presence of AgI species and the small stars (*) at 2θ = 44.4° indicate the presence of Ag0.
Figure 1. Powder XRD patterns of pristine TiO2 and Ag/TiO2 materials. Small squares (□) at 2θ = 33.2° indicate the presence of AgI species and the small stars (*) at 2θ = 44.4° indicate the presence of Ag0.
Catalysts 09 00323 g001
Figure 2. (a) UV-Visible spectra of TiO2 and Ag/TiO2 as pressed disks. Inset shows the material absorption magnification at 400–700 n; (b) surface plasmon resonance (SPR) effect of Ag species deposited on TiO2 surface, modified from reference [48].
Figure 2. (a) UV-Visible spectra of TiO2 and Ag/TiO2 as pressed disks. Inset shows the material absorption magnification at 400–700 n; (b) surface plasmon resonance (SPR) effect of Ag species deposited on TiO2 surface, modified from reference [48].
Catalysts 09 00323 g002
Figure 3. (a) Diffuse reflectance spectra and (b) band gap calculation of TiO2 and Ag/TiO2X% (X = 1–4).
Figure 3. (a) Diffuse reflectance spectra and (b) band gap calculation of TiO2 and Ag/TiO2X% (X = 1–4).
Catalysts 09 00323 g003
Figure 4. X-ray photoelectron (XP) spectra of Ag/TiO2 4%: (a) Survey scan spectrum and peak fitting results of (b) Ag 3d, (c) Ti 2p, and (d) O 1s.
Figure 4. X-ray photoelectron (XP) spectra of Ag/TiO2 4%: (a) Survey scan spectrum and peak fitting results of (b) Ag 3d, (c) Ti 2p, and (d) O 1s.
Catalysts 09 00323 g004
Figure 5. SEM images of Ag/TiO2 4% in (a) 50,000 and (b) 100,000 magnifications (backscaterred electron (BSE) detector used for images (c) and (d)).
Figure 5. SEM images of Ag/TiO2 4% in (a) 50,000 and (b) 100,000 magnifications (backscaterred electron (BSE) detector used for images (c) and (d)).
Catalysts 09 00323 g005
Figure 6. TEM images of Ag/TiO2 2% (a,b) and TEM images of Ag/TiO2 4% (ce).
Figure 6. TEM images of Ag/TiO2 2% (a,b) and TEM images of Ag/TiO2 4% (ce).
Catalysts 09 00323 g006
Figure 7. CH4 concentration over time for the Ag/TiO2 2% sample in batch cleaning, and CO2 reduction with Hg/Xe lamp and LED, as well as CO2 reduction with metallic state Ag0/TiO2 2%. Bare TiO2 is used as a reference.
Figure 7. CH4 concentration over time for the Ag/TiO2 2% sample in batch cleaning, and CO2 reduction with Hg/Xe lamp and LED, as well as CO2 reduction with metallic state Ag0/TiO2 2%. Bare TiO2 is used as a reference.
Catalysts 09 00323 g007

Share and Cite

MDPI and ACS Style

Handoko, C.T.; Moustakas, N.G.; Peppel, T.; Springer, A.; Oropeza, F.E.; Huda, A.; Bustan, M.D.; Yudono, B.; Gulo, F.; Strunk, J. Characterization and Effect of Ag(0) vs. Ag(I) Species and Their Localized Plasmon Resonance on Photochemically Inactive TiO2. Catalysts 2019, 9, 323. https://doi.org/10.3390/catal9040323

AMA Style

Handoko CT, Moustakas NG, Peppel T, Springer A, Oropeza FE, Huda A, Bustan MD, Yudono B, Gulo F, Strunk J. Characterization and Effect of Ag(0) vs. Ag(I) Species and Their Localized Plasmon Resonance on Photochemically Inactive TiO2. Catalysts. 2019; 9(4):323. https://doi.org/10.3390/catal9040323

Chicago/Turabian Style

Handoko, Chanel Tri, Nikolaos G. Moustakas, Tim Peppel, Armin Springer, Freddy E. Oropeza, Adri Huda, Muhammad Djoni Bustan, Bambang Yudono, Fakhili Gulo, and Jennifer Strunk. 2019. "Characterization and Effect of Ag(0) vs. Ag(I) Species and Their Localized Plasmon Resonance on Photochemically Inactive TiO2" Catalysts 9, no. 4: 323. https://doi.org/10.3390/catal9040323

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