Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Preparation and SERS applications of Ta2O5 composite nanostructures

Open Access Open Access

Abstract

Noble metal and semiconductor composite substrates possess high sensitivity, excellent stability, good biocompatibility, and selective enhancement, making them an important research direction in the field of surface-enhanced Raman scattering (SERS). Ta2O5, as a semiconductor material with high thermal stability, corrosion resistance, outstanding optical properties, and catalytic performance, has great potential in SERS research. This study aims to design and fabricate a composite SERS substrate based on Ta2O5 nanostructures, achieving optimal detection performance by combining the urchin-like structure of Ta2O5 with silver nanoparticles (Ag NPs). The urchin-like Ta2O5 nanostructures were prepared using a hydrothermal reaction method. The bandgap was modulated through structure design and the self-doping technique, the charge transfer efficiency and surface plasmon resonance effects were improved, thereby achieving better SERS performance. The composite substrate enables highly sensitive quantitative detection. This composite SERS substrate combines the electromagnetic enhancement mechanism (EM) and chemical enhancement mechanism (CM), achieving ultra-low detection limits of 10−13 M for R6G. Within the concentration range above 10−12 M, there is a good linear relationship between concentration and peak intensity, demonstrating excellent quantitative analysis capabilities. Furthermore, this composite SERS substrate is capable of precise detection of analytes such as crystal violet (CV) and methylene blue (MB), holding broad application prospects in areas such as food safety and environmental monitoring.

© 2023 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Raman scattering spectroscopy is a non-elastic scattering spectroscopic analysis method based on the energy band structure of materials. Due to the different molecular structures of different substances, their energy band structures vary, resulting in distinct Raman spectra. Consequently, Raman spectroscopy is widely employed for identification and analysis of material composition and molecular structure, becoming an efficient analytical tool [1]. However, the Raman signals of the target substances themselves are typically extremely weak, necessitating the utilization of SERS technology to amplify the Raman signals of the target substances [2]. SERS technology employs special substrate materials and structures that can significantly enhance the Raman signals of the target molecules [3], with enhancement factors often exceeding 1012. This technique enables the detection of target substances at low concentrations and can even achieve single-molecule-level detection [4].Currently, SERS technology has been widely applied in various fields such as chemical analysis, medical diagnostics, food safety and environmental monitoring, demonstrating significant application potential and promising prospects. With continuous technological advancements, SERS technology will continue to drive the development of molecular structure analysis, providing us with more precise and efficient analytical tools [5].

The recognized mechanisms of SERS mainly include EM and CM [6]. In the SERS effect, EM plays a primary role, but CM also exhibits unique and specific enhancement effects. The synergistic interaction of these two mechanisms can achieve superior enhancement effects and reach lower detection limits [7].EM primarily achieves enhancement effects through localized surface plasmon resonance (LSPR) [8], which has a wide range of applications and can enhance the Raman signals of various molecules. In the EM mechanism, the incident laser generates a plasma effect on the surface, which absorbs more of the incident light, reduces reflection, strengthens scattering, and thus enhances the Raman effect [9]. However, this mechanism exhibits some limitations in terms of specificity. Meanwhile, CM achieves enhancement effects through charge transfer efficiency between the analyte and substrate [10]. CM mainly uses charge transfer to make it easier for electrons to enter the conduction band from the valence band of the measured substance, thereby allowing more electrons to transition back from the conduction band and generate Raman effects. This mechanism provides higher specificity enhancement, facilitating the detection and analysis of specific molecules. By combining EM with CM, better SERS effects can be achieved. EM offers broad applicability and high gain, while CM imparts specificity enhancement, leading to higher gains for specific molecules. This integrated approach not only increases the intensity of Raman signals but also further reduces the detection limit, enabling SERS technology to unleash its full potential in detecting substances at lower concentrations.

Noble metal materials were initially widely recognized for their application in SERS effects, particularly due to the outstanding SERS performance of noble metal nanostructures, such as Au/Ag NPs [11]. With deeper research into noble metal nanostructures and SERS mechanisms, researchers have been continuously exploring the application of other materials in SERS. For example, organic dyes [12], semiconductor nanostructures [13], and two-dimensional materials have also been utilized for SERS [14], and have shown certain enhancement effects. While exploring materials, the nanostructure of the substrate is also constantly being studied and improved, such as Pyramid structure [15], nanorod array [16], and nanotubes [17]. In particular, the application of semiconductor materials in the Raman field has experienced rapid development. Semiconductor materials have become highly promising SERS substrate materials due to their advantages of high chemical stability, good biocompatibility, high carrier mobility [18], and controllable preparation processes [19,20]. Semiconductor nanomaterials such as ZnO, ZnS, Pb3O4, CuO, and TiO2 have demonstrated strong SERS signals, making them widely regarded as excellent SERS substrate materials [21].

In nanostructured semiconductor materials, the SERS mechanisms can be categorized into three cases: Firstly, charge transfer efficiency is one of the main sources of the SERS effect in semiconductors and belongs to the chemical enhancement mechanism. This enhancement effect is closely related to the energy level matching between the semiconductor substrate and the target molecules. Charge transfer efficiency occurs when the highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO) of the target molecule align with the conduction and valence band levels of the semiconductor. This process alters the polarizability tensor and electron density distribution of the molecule, resulting in a pronounced SERS effect. Secondly, when light is incident on the surface of semiconductor nanostructures, it can induce LSPR, leading to the generation of the SERS effect, which operates on a mechanism similar to that of metal materials. The third mechanism occurs when the dimensions of the nanostructures or cavities on the substrate are equal to the wavelength of the incident light, resulting in Mie scattering, which further improves the SERS performance of the substrate [22].One significant advantage of using semiconductors as Raman substrates is the ability to modify properties such as HOMO, LUMO, and surface charge density through various means, including adjusting nanostructure, doping, and introducing defects. By modifying these properties, the charge transfer efficiency and LSPR effects can be enhanced, leading to improved Raman enhancement and specificity [13].

While there have been numerous studies on the SERS effect of oxide semiconductors substrates, relatively few studies have focused on SERS performance of tantalum-based oxide materials (Ta2O5). However, tantalum-based oxide materials possess exceptional thermal stability and corrosion resistance, along with excellent optical and catalytic properties, thus holding significant potential in the field of SERS [23].Although there is relatively limited research on the SERS performance of tantalum-based oxide materials, existing studies have shown their potential in SERS applications. By manipulating the structure and surface properties of tantalum-based oxide materials, such as through nanoscale engineering, surface modification, or doping, their SERS performance can be further optimized to achieve higher sensitivity and specific enhancement [24,25]. Additionally, tantalum-based oxide materials can be combined with other functional materials to expand their applications in the field of SERS.

The application of Ta2O5 materials in the SERS field faces challenges such as relatively weak enhancement and the difficulty of achieving charge transfer efficiency due to its wide-bandgap semiconductor nature. To address these issues, this study focused on optimizing the SERS performance of Ta2O5 substrates through morphology control, band engineering, and material composites. In terms of morphology control, the preparation of urchin-like nano structures in Ta2O5 effectively increased the surface area and the number of active sites, thereby enhancing the SERS effect. In terms of band engineering, the self-doping effect was achieved in Ta2O5 structures prepared through hydrothermal reactions, resulting in a change in the charge density and band structure of Ta2O5, thereby regulating its Raman effect [26]. Finally, in the aspect of material composites, a composite structure was formed by combining Ta2O5 with Ag NPs that exhibit strong SERS activity. The SERS effect of Ag NPs was utilized, and additional modification sites were provided on the Ta2O5 surface, leading to synergistic enhancement.

In this study, a Ag NPs/Ta2O5 composite substrate was designed and prepared. With the synergistic enhancement of CM and EM, Three low concentration substances were detected, namely Rhodamine 6 G (R6G), crystal violet (CV), and methylene blue (MB), with detection limits of 10−13 M, 10−11 M, and 10−11 M, respectively. Moreover, the substrate exhibited good uniformity and strong stability, allowing for substrate recycling.

2. Experimental

2.1 Materials

Tantalum tablets (Ta, 99.9%), hydrofluoric acid (HF, AR, ≥ 40%), hydrogen peroxide solution (H2O2, AR, 30 wt. % in H2O), silver nitrate (AgNO3, AR, 99.8%) and polyvinylpyrrolidone (PVP, average Mw 10000) were purchased from Aladdin Co., Ltd. (Shanghai, China). Acetone (C3H6O, AR, ≥ 99.5%), ethylene glycol (C2H6O2, AR, ≥ 99.5%) and ethanol absolute (C2H6O, AR, ≥ 99.7%) were bought from Sinopharm Chemical Reagent Co., Ltd. Deionized water is prepared by laboratory equipment.

2.2 Synthesis of Ag NPs and Ta2O5 nanostructures

Synthesis of Ag NPs was achieved through a chemical method [27]. Firstly, in a 20 mL round-bottom flask, an ethylene glycol solution was added and placed on a temperature-controlled magnetic stirrer. When the temperature reached 70 °C, 0.25 g PVP was added to the flask, followed by further heating to 135 °C. At this temperature, 0.05 g of AgNO3 was added. The mixture was heated for 3 hours until it appeared milky white, after which the liquid was removed and allowed to cool. After cooling, 45 mL of acetone was added to the mixture for cleaning and centrifugation. After several cycles of cleaning and centrifugation, a solution of Ag NPs was obtained.

The Ta2O5 nanostructures were synthesized using a hydrothermal reaction method [28]. Firstly, a tantalum sheet with dimensions of 10 mm × 10 mm and a thickness of 1 mm was taken and subjected to grinding, polishing, and cleaning treatments. The treated tantalum sheet was then placed in a reaction vessel. Next, 5 mL of 30% mass concentration H2O2 and 1 mL of 40% mass concentration HF were injected into the reaction vessel. The reaction vessel was placed in a drying oven, and the oven temperature was set to 220 °C with a reaction time of 24 h. After the reaction was completed, the tantalum sheet was taken out and subjected to drying treatment. Multiple layers of Ta2O5 structures were observed on the surface of the tantalum sheet.

2.3 Preparation of the composite substrate

The preparation process of the composite substrate is shown in Fig. 1. The Ag NPs solution was subjected to 30 minutes of ultrasonic treatment to achieve uniform dispersion of the Ag NPs. Next, the Ta2O5 sample prepared by the hydrothermal reaction method was immersed in the Ag NPs solution and left to soak for 20 minutes. Subsequently, the sample was placed in a drying oven at a temperature of 60 °C Celsius for continuous drying for 24 h to obtain the composite substrate sample.

 figure: Fig. 1.

Fig. 1. Preparation process of Ag NPs/ Ta2O5 composite SERS substrate.

Download Full Size | PDF

2.4 Experimental instrument

Scanning electron microscopy (SEM, ZEISS Gemini Sigma 500) was employed to characterize the morphology and structure of Ag NPs, urchin-like Ta2O5 and the composite substrate. Energy-dispersive spectroscopy (EDS, Bruker QUANTAX EDS) was used to characterize the elemental composition of the composite substrate. The ultraviolet-visible spectrophotometer (UV-Vis, Alpha-1500) was used to measure the absorption spectrum of urchin-like Ta2O5 and determine the bandgap width. Ultraviolet photoelectron spectroscopy (UPS, Thermo Scientific Escalab Xi+) with an electron energy of 21.22 eV was utilized to determine the valence band position and Fermi level of the sample. X-ray photoelectron spectroscopy (XPS, Thermo Scientific Escalab 250Xi) was employed to characterize the surface chemical composition and chemical states of Ta2O5. X-ray diffraction (XRD, Rigaku D/MAX-RB) was used to determine the crystal structure of urchin-like Ta2O5. Raman spectroscopy (Horiba HR Evolution 800) with a laser wavelength of 532 nm was employed to investigate the Raman properties of the composite substrate. Fifteen random locations were selected for Raman spectroscopy, and the average values were calculated.

3. Results and discussion

3.1 Characterization of Ta2O5 nanostructures and composite substrates

In the experiment, the Ag NPs were prepared using the afore-mentioned chemical method. In this method, ethylene glycol was used as a reducing agent to reduce AgNO3 to Ag NPs at a temperature of 135 °C. Polyvinylpyrrolidone (PVP) was used as a capping agent to promote the formation of uniform Ag NPs structures [29]. From SEM images, the prepared Ag NPs showed a relatively uniform diameter distribution, mainly ranging from 80 nm to 120 nm, and exhibited good consistency (Fig. 2(a)). Incorporating these Ag NPs into the composite substrate would contribute to achieving significant Raman enhancement effects.

 figure: Fig. 2.

Fig. 2. (a-b) SEM image of Ag NPs, urchin-like Ta2O5. (c-d) SEM image of Ag NPs/Ta2O5 composite substrate. (e-f) TEM and HRTEM image of urchin-like Ta2O5. (g) electron diffraction pattern of urchin-like Ta2O5.

Download Full Size | PDF

The urchin-like Ta2O5 nanostructures were prepared using the hydrothermal method described in 2.2. In this process, H2O2 and HF were used to corrode the tantalum sheet, resulting in the formation of a layer of urchin-like Ta2O5 nanostructures on the tantalum surface. Due to the presence of a dense Ta2O5 protective layer on the tantalum surface, only a few acids can penetrate this barrier and corrode the tantalum sheet. Nitric acid, sulfuric acid, or even aqua regia, which are highly corrosive, cannot penetrate this protective layer. However, HF is an exception. The fluoride ions in HF are small enough to penetrate the oxide layer and react with the tantalum metal, forming a complex structure. Although HF is a weak acid, it is one of the few inorganic agents that can react with the tantalum sheet. The addition of H2O2 enhances the oxidizing properties of HF and promotes the formation of Ta2O5 nanostructures. The specific reaction equation is as follows [28]:

$$2\textrm{Ta}\; + {\; }14\textrm{HF} \to 2{\textrm{H}_2}\textrm{Ta}{\textrm{F}_7}{\; } + {\; }5{\textrm{H}_2} \uparrow {\; }$$
$$2{\textrm{H}_2}\textrm{Ta}{\textrm{F}_7}{\; } + {\; }5{\textrm{H}_2}\textrm{O} \to \; \textrm{T}{\textrm{a}_2}{\textrm{O}_5} \downarrow + {\; }14\textrm{HF}$$
$${\textrm{H}_2}{\textrm{O}_2}{\; } + {\; }{\textrm{H}_2} \to {\; }2{\textrm{H}_2}\textrm{O}$$

The Ta2O5 nanostructures formed by the hydrothermal reaction exhibit a multi-layered structure on a macroscopic scale. SEM images show that each layer is composed of Ta2O5 nanostructures with a sea urchin-like morphology (Fig. 2(b)). The nano-spikes of these urchin structures have a length of approximately 400 nm, while the base diameter is around 100 nm. The size of these nanostructures shows a relatively uniform distribution overall and exhibits good consistency.

Due to the unique structural features of the prepared Ta2O5, it is only necessary to immerse it in the Ag NPs solution, and the Ag NPs will spontaneously fill the gaps between the Ta2O5 nano-spikes, forming a tightly bonded Ag NPs/Ta2O5 composite substrate (Fig. 2(c), Fig. 2(d)). This processing method is relatively simple and does not require complex fabrication steps. The good bonding between the Ag NPs and the Ta2O5 nanostructures allows the composite substrate to exhibit highly consistent structural characteristics.

To further investigate the properties of Ta2O5 in urchin-like structures, the lattice structure of post-treated Ta2O5 nanospikes was characterized using transmission electron microscopy (TEM) and high-resolution transmission electron microscopy (HRTEM). The HRTEM image (Fig. 2(f)) clearly shows the lattice spacing of Ta2O5 in the urchin-like structure. The hydrothermally synthesized Ta2O5 exhibits a single crystal structure predominantly growing along the (010) direction, with a lattice plane spacing of 0.385 nm, corresponding to the (010) crystal plane of orthorhombic Ta2O5 (JCPDS 79-1375) [30,31].

The results of EDS testing show that the distribution of oxygen and tantalum elements in the nanostructured Ta2O5 is extremely uniform (Fig. 3(a)), with a ratio close to 2:1 (Fig. 3(b)). Although this ratio does not exactly match the chemical formula of Ta2O5, subsequent XPS analysis reveals that this is due to the presence of a significant amount of oxygen vacancies in the nanostructured Ta2O5 after the hydrothermal reaction. Additionally, the use of hydrofluoric acid during the hydrothermal reaction results in a small amount of residual fluorine on the surface of the nanostructured Ta2O5. However, further X-ray diffraction (XRD) analysis confirms that these fluorine elements do not remain in the lattice. Furthermore, the content of fluorine is very low and does not significantly affect the properties of the sample.

 figure: Fig. 3.

Fig. 3. (a-b) EDS image of urchin-like Ta2O5. (c) UV-Vis patterns of Ag NP, R6G and R6G on Ag NPs/Ta2O5. (d) XRD patterns of urchin-like Ta2O5. (e-f) UV-Vis patterns of urchin-like Ta2O5 showing wavelength and energy.

Download Full Size | PDF

XRD analysis of the hydrothermally prepared urchin-like Ta2O5 sample revealed XRD patterns (Fig. 3(d)) corresponding to the lattice constants of orthorhombic Ta2O5 (JCPDS 79-1375) phase, which is consistent with the TEM results, confirming the prepared tantalum oxide as Ta2O5.Apart from these characteristic peaks, no other significant diffraction peaks were observed, indicating that no other by-products such as H2TaO6 and Ta2O5·nH2O were formed during the hydrothermal reaction, or if present, they were present in only trace amounts. This also suggests that the fluorine element in hydrofluoric acid is only present on the surface of Ta2O5 nanospikes and hydrofluoric acid acts solely as an inorganic agent in the reaction, used for tantalum sheet etching, promoting the crystallization of Ta2O5 and the formation of the urchin-like structure, without being incorporated into the generated Ta2O5 lattice. The strongest peak in the XRD pattern corresponds to the (010) crystal plane, indicating a preferential growth along the b-axis, perpendicular to the (010) crystal plane. Other crystal planes such as (002), (012), (020), (2410) exhibit weaker crystallinity, showing smaller and broader peaks, consistent with the TEM results.

To measure the bandgap width and relative changes of the urchin-like Ta2O5 nanostructures, UV-Vis spectrophotometry was conducted on both purchased Ta2O5 powder and prepared Ta2O5 nanostructures in the wavelength range of 200 nm to 800 nm. As shown in Fig. 3(e) and Fig. 3(f), compared to the purchased Ta2O5 powder, the urchin-like Ta2O5 prepared through the hydrothermal reaction broadens the absorption frequency band, indicating a narrower bandgap width. By converting the energy-absorption spectra, an approximate bandgap width of 3.84 eV was calculated for the purchased Ta2O5 powder, while the bandgap width of the urchin-like Ta2O5 was approximately 2.78 eV, consistent with reported data on other hydrothermally synthesized Ta2O5 nanostructures [32]. The reduction in bandgap width is attributed to the excess tantalum sheet used in the experiment, which leads to the partial reduction of Ta5+ to Ta4+. The formation of Ta4+ is accompanied by the formation of oxygen vacancies. The presence of oxygen vacancies introduces defect energy levels, altering the overall band structure and significantly reducing the bandgap width [33,34]. Thus, it can be observed that the urchin-like Ta2O5 exhibits a significantly reduced bandgap width, thereby extending its absorption range for laser excitation, which positively contributes to achieving charge transfer efficiency and enhancing SERS performance of the substrate.

From Fig. 3(c), it can be seen that Ag NPs have a very wide spectral response, with absorption peaks around 450 nm. The absorption peak of R6G ranges from 500 to 550 nm. After dropping R6G onto the Ag NPs/Ta2O5 composite substrate, the absorption peak appears at around 500 nm. Due to the fact that the Raman spectrometer in the laboratory only has two types of lasers: 325 nm and 532 nm, a 532 nm laser was selected as the testing light source for this experiment, which is also the most common and effective Raman testing light source.

In order to gain a deeper understanding of the composition and band information of urchin-like Ta2O5, XPS characterization was performed on the purchased Ta2O5 powder and the prepared urchin-like Ta2O5. As shown in Fig. 4(a) and Fig. 4(d), both exhibit characteristic double peaks belonging to Ta4f. Notably, in the urchin-like Ta2O5, the characteristic peaks of Ta4f are observed to have a blue shift, indicating a higher binding energy. During the fitting process, it was observed that the Ta4f spectrum of urchin-like Ta2O5 exhibits, in addition to the double peaks at 26.3 eV (Ta5 + 4f5/2) and 28.2 eV (Ta5 + 4f7/2) attributed to Ta5+, two distinct double peaks located at 25.9 eV and 27.8 eV. The difference between these two peaks is similar to that of Ta5+, and their shapes are analogous. Through literature review, coupled with results from EDS and UV tests, it was confirmed that these double peaks are attributed to Ta4+ and represent Ta4 + 4f5/2 and Ta4 + 4f7/2. This finding indicates the presence of Ta4+ within the urchin-like Ta2O5 and further suggests the existence of oxygen vacancies [25].

 figure: Fig. 4.

Fig. 4. (a-b) XPS spectra of Ta4f, O1s of urchin-like Ta2O5. (c) XPS survey spectrum of Ag NPs/ Ta2O5 composite substrate. (d-e) XPS spectra of Ta4f, O1s of purchased Ta2O5. (f) Valence-band XPS spectra of purchased Ta2O5 and urchin-like Ta2O5. (g-h) UPS spectra of urchin-like Ta2O5.

Download Full Size | PDF

Analysis of the O1s XPS spectra (Fig. 4(b), Fig. 4(e)) revealed that the purchased Ta2O5 powder exhibited two peaks located at 529.9 eV and 531.6 eV, indicating the presence of oxygen in two forms corresponding to lattice oxygen and hydroxyl groups bound to the surface. In contrast, the urchin-like Ta2O5 displayed three peaks at 530.2 eV, 531.3 eV, and 532.4 eV, indicating the presence of oxygen in three forms: lattice oxygen, oxygen adsorbed by oxygen vacancies, and hydroxyl groups [35,36]. This indicates that the urchin-like Ta2O5 prepared through the hydrothermal reaction has a significant amount of oxygen vacancies and a certain amount of hydroxyl groups. The presence of oxygen vacancies and the unique nanostructure significantly influence the band structure and properties of Ta2O5, facilitating charge transfer efficiency and achieving enhanced Raman effects.

To accurately determine the position of the energy bands, UPS testing was conducted on the urchin-like Ta2O5 prepared through the hydrothermal reaction. The excitation energy of the instrument was set to 21.22 eV. According to the UPS spectrum (Fig. 4(g), relative to the vacuum level, the valence band maximum of the urchin-like Ta2O5 was located at -6.66 eV. Based on the UV testing results, it can be inferred that the conduction band minimum is located at -3.88 eV. According to the UPS spectrum with a bias voltage of -5 V (Fig. 4(h)), the work function of the urchin-like Ta2O5 was determined to be 3.75 eV (21.22 eV-17.47 eV), indicating that the Fermi level is at -3.75 eV. Analysis of the band structure reveals that the prepared urchin-like Ta2O5 has a Fermi level higher than the conduction band minimum, indicating a degenerate semiconductor. This is due to the presence of a significant amount of oxygen vacancies in the urchin-like Ta2O5 structure prepared through the hydrothermal reaction, which results in self-doping, increases the carrier concentration, provides more excited states, and further enhances the interaction with adsorbed molecules, facilitating charge transfer efficiency and enhancing the Raman signal of the analyte molecules [33].

3.2 Raman results

In this study, the Raman tests used a laser wavelength of 532 nm. Except for the test in Fig. 5(a), which used a laser power of 24 µW, the laser power used in other Raman tests was 4.8 µW. The enhancement effect of tantalum sheet and commercial tantalum oxide in Fig. 5(a) is very weak, it requires the use of higher power lasers to obtain Raman signals. If higher power lasers are also used for other tests, there will be a phenomenon of peak shaving in the spectrum at high concentrations of R6G, so a lower power laser was chosen. A widely used SERS probe molecule, R6G, was chosen for conducting comprehensive comparative tests on the SERS performance of tantalum sheet, commercial Ta2O5, urchin-like Ta2O5 structures, Ag NPs, and Ag NPs/Ta2O5 composite substrates. The results (Fig. 5(a)) showed that the Ag NPs/Ta2O5 composite substrate exhibited the strongest Raman enhancement, with higher peak intensity compared to the substrate with Ag NPs alone, and achieved an extremely low detection limit of 10−13 M (Fig. 5(c)) with an enhancement factor of up to 1010. This significant performance improvement is mainly attributed to the LSPR effect induced by Ag NPs and the energy level matching between the urchin-like Ta2O5 structures and R6G, promoting charge transfer efficiency. Furthermore, a comparative study revealed that the nanostructured Ta2O5 prepared through the hydrothermal reaction exhibited significantly better SERS performance than the commercially purchased Ta2O5 substrate. On one hand, the unique structure of nanostructured Ta2O5 can generate a stronger LSPR effect, which was verified by subsequent Comsol simulations. On the other hand, the nanostructured Ta2O5 structure has a larger specific surface area, and XPS results indicated the presence of a large number of oxygen vacancies on its surface. These characteristics allow the nanostructured Ta2O5 to modify the band structure and facilitate charge transfer efficiency.

 figure: Fig. 5.

Fig. 5. (a) The Raman spectra of R6G on Ta sheet, purchased Ta2O5, urchin-like Ta2O5, Ag NPs and Ag NPs/Ta2O5 composite substrate using a laser power of 24 µW. (b) Raman spectra of R6G on the Ag NPs/Ta2O5 composite substrates from 10−7 M to 10−12 M using a laser power of 4.8 µW. (c) Ag NPs/Ta2O5 composite substrate's lowest detection limit for R6G. (d) The relationship between R6G concentration and the intensities of the Raman peaks at 611 cm-1 and 1362 cm-1. (e) Raman mapping image of R6G on the composite substrate. (f-g) Raman spectra of CV and MB on Ag NPs/Ta2O5 composite substrate at concentrations ranging from 10−7 M to 10−12 M.

Download Full Size | PDF

Additionally, the relationship between the Raman signal intensity of the composite substrate and the concentration of R6G was tested (Fig. 5(d)). R6G was drop-cast onto the composite substrate in volumes of 2 µL over a concentration range of 10−12 to 10−7 M, followed by Raman measurements. The test results showed a linear relationship between the logarithm of the intensity at 611 cm-1 and 1362 cm-1 peaks and the logarithm of the concentration, with correlation coefficients (R2) of 0.98162 and 0.98969, respectively. This provides a basis for quantitative measurement of the concentration of the analyte on the substrate to some extent. Raman mapping (Fig. 5(e)) revealed that the intensity of the 611 cm-1 peak was relatively uniform within the tested range, indicating good uniformity of the substrate. This is mainly attributed to the urchin-like Ta2O5 structures tightly covering the entire surface of the substrate, with Ag NPs uniformly embedded in the gaps of the nano spikes. Therefore, the substrate exhibits good consistency and uniformity, resulting in good uniformity in Raman measurements.

By comparing and testing the SERS performance of tantalum foil, commercial Ta2O5, nanostructured Ta2O5, Ag NPs, and Ag NPs/Ta2O5 composite substrates, this study found that the Ag NPs/Ta2O5 composite substrate exhibited the highest peak intensity, superior to the use of Ag nanoparticles alone, and achieved a low detection limit of 10−13 M. The nanostructured Ta2O5 exhibited significantly better SERS performance than the commercial Ta2O5 substrate, attributed to its stronger LSPR effect and larger surface area, as well as the modification of the band structure by the presence of oxygen vacancies, facilitating charge transfer efficiency. Additionally, through testing the Raman signal intensity and R6G concentration of the composite substrate, quantitative measurement of the concentration of the analyte on the substrate can be achieved. Raman mapping results demonstrated good uniformity of the composite substrate, which is attributed to the urchin-like Ta2O5 structures covering the entire surface and the uniform embedding of Ag NPs in the gaps of the nano spikes, ensuring the consistency and uniformity of the substrate.

To validate the universality of the substrate enhancement effect, two dye molecules, CV (Fig. 5(f)) and MB (Fig. 5(g)), were tested. The test results showed that both CV and MB molecules exhibited clear characteristic peaks even at a concentration of 10−11 M, indicating a strong enhancement effect of the substrate on these two dye molecules.

By calculating RSD parameters, the uniformity and consistency of the substrate can be characterized. Prepare 6 Ag NPs/Ta2O5 composite substrates, use 10−10 M R6G as probes, select 5 test points for each substrate, measure 30 sets of spectral data as seen in Fig. 5(d), and calculate the RSD value of 611 cm-1 peak. The formula is as follows:

$$RSD = \frac{S}{{\bar{I}}} \times 100\%= \frac{{\sqrt {\mathop \sum \nolimits_{i = 1}^n {{({{I_i} - \bar{I}} )}^2}} }}{I}$$

The SERS signal strength of R6G 611 cm-1 peak is represented by I, and $\bar{I}$ represents the average value of all 611 cm-1 peaks at the test points. The final measured substrate RSD value is about 9.16%, indicating that the Ag NPs/Ta2O5 substrate has good uniformity and reproducibility.

To quantify the sensitivity of the substrate, R6G was used as the probe and silica substrate as the benchmark. Calculate the EF for R6G using the following formula:

$$EF = \frac{{{I_{SERS}}/{N_{SERS}}}}{{{I_{Si{O_2}}}/{N_{Si{O_2}}}}}$$

In the formula, ${I_{SERS}}$ and ${I_{Si{O_2}}}$ are the intensity of the SERS peak and the normal Raman intensity on SiO2 substrate respectively, and ${N_{SERS}}$ and ${N_{Si{O_2}}}$ are the number of R6G molecules per unit volume of SERS and normal Raman scattering. As shown in Fig. 6(a), the peak strength of 10−2 M R6G at 611 cm−1 on SiO2 is 836 and the peak intensity of 10−12 M R6G on Ag NPs/Ta2O5 composite substrate is 1624. As a result, the average EF of Ag NPs/Ta2O5 composite substrate is 1.94 × 1010, which is higher than that of several Ta2O5-based or Ag NPs-based composite SERS substrates that have been reported, as shown in Table 1. As a result, the average EF of Ag NPs/Ta2O5 composite substrate was 1.94 × 1010, which is very excellent among several reported composite SERS substrates based on Ta2O5 or Ag NPs, as shown in Table 1.

 figure: Fig. 6.

Fig. 6. (a) The Raman spectra of 10−2 M R6G on SiO2 and 10−12 M R6G on Ag NPs/Ta2O5. (b-c) The results of the 30-day stability test of the composite substrate. (d) SEM image of Ag NPs/Ta2O5 substrate after HNO3 cleaning. (e) Comparison of Raman spectra before and after recycling treatment.

Download Full Size | PDF

Tables Icon

Table 1. Sensitivity Comparison of Several Ta2O5-based or Ag NPs-based Composite SERS Substrates

To verify the stability of the substrate, stability tests were conducted for a period of 30 days. After preparing the composite substrate, it was stored in the atmosphere at a temperature of around 25 °C at room temperature. Every 5 days, Raman spectra were measured using R6G samples at a concentration of 10−10 M, and the results are shown in Fig. 6(b). The intensity of the 611 cm-1 peak was chosen for statistical analysis (Fig. 6(c)). In the initial 10 days, the peak intensity decayed rapidly and then stabilized. At the end of the 30-day test, the Raman signal remained above 80% of the initial peak intensity. This indicates that the urchin-like Ta2O5 structure in the composite substrate is highly stable, unaffected by oxidation, and exhibits excellent corrosion resistance and high-temperature stability, making it suitable for harsh detection environments, which is of great significance in practical applications. The main reason for the substrate decay is the oxidation of Ag NPs; however, due to the close coupling between Ag NPs and the urchin structure, it provides a certain level of protection, slowing down the oxidation process of Ag NPs. After the oxidation of Ag NPs, the substrate can be dissolved by immersing it in HNO3, removing the oxidized Ag NPs. Since Ta2O5 has excellent acid corrosion resistance, the urchin-like Ta2O5 structure is not damaged (Fig. 6(d)). After ultrasonic cleaning with alcohol and deionized water, Ag NPs can be re-added to form a new composite substrate, achieving the reusability of the substrate. The Raman EF of the substrate produced after recycling can be restored (Fig. 6(e)).

3.3 Comsol simulation

To gain a deeper understanding of the electromagnetic enhancement mechanism of the composite substrate and assist in the experimental analysis, COMSOL software was used for simulation to study the electric field distribution of various structures under the excitation of a 532 nm laser [39]. Through these simulations, it is possible to explore the electromagnetic enhancement mechanism more comprehensively and provide guidance for optimizing the nanostructure of the substrate. Corresponding simulation parameters were set based on the actual situation of the experimental device and SEM images. The incident laser wavelength was set to 532 nm, and the diameter of the Ag NPs was 100 nm. The Ta2O5 nanostructure was simplified to a conical shape with a base diameter of 100 nm and a height of 400 nm, resembling the morphology of a urchin. By simulating this structure, corresponding electric field distribution images were obtained, as shown in the Fig. 7.

 figure: Fig. 7.

Fig. 7. Simulated electric field distribution of Ta2O5, Ag NPs, urchin-like Ta2O5 and Ag NPs/Ta2O5 composite substrates.

Download Full Size | PDF

From the results in Fig. 7(a), it can be observed that compared to Ta2O5 nanoparticles, the Ta2O5 nano-spikes can generate a stronger, more extensive, and more uniformly distributed excitation electric field. As a result, the Ta2O5 nano-spikes exhibit a stronger LSPR effect, leading to better enhancement. In Fig. 7(b), a significant LSPR effect can be observed between Ag NPs, leading to a strong local electric field. Figure 7(c) shows the simulation of a single Ta2O5 sea urchin-like nanostructure, revealing a stronger localized electric field at the base of the spikes, which is enhanced compared to individual nano-spikes but weaker than that between the Ag NPs. Figure 7(d) presents the simulation of the Ag NPs/Ta2O5 composite substrate structure. In this structure, there exists a strong and extensive excitation electric field between the Ag nanoparticles and the Ta2O5 nano-spikes, with the intensity even surpassing that between the Ag NPs. The exceptionally strong excitation field can be attributed not only to the unique sea urchin-like structure that tightly binds the Ag NPs and Ta2O5 nano-spikes but also, primarily, to the achievement of energy level alignment between the Ag NPs and Ta2O5 nano-spikes, facilitating charge transfer efficiency (CM). It can be concluded that the LSPR effect of composite substrates is very strong and superior to pure Ag NPs substrates, exhibiting a wide distribution range. Consequently, the Ag NPs/Ta2O5 composite substrate demonstrates excellent electromagnetic enhancement.

3.4 Theoretical analysis

The enhanced performance of the Ag NPs/Ta2O5 composite substrate is a result of EM and CM. The EM enhancement is primarily attributed to the LSPR effect of the Ag nanoparticles. The CM enhancement arises from the energy level alignment between R6G molecules, Ta2O5, and Ag NPs. The synergistic interplay between the EM and CM mechanisms significantly enhances the substrate's SERS effect, enabling highly sensitive detection and analysis of molecules.

Based on the EM mechanism, employing Comsol electromagnetic simulation technology allows us to draw the following conclusions: LSPR effects exist between the Ag nanoparticles and the Ta2O5 nanocones, as well as among the Ag nanoparticles themselves, resulting in strong localized electric fields and subsequent electromagnetic enhancement. In this process, a substantial portion of the incident light energy is absorbed by the surface plasmon waves, leading to a sharp decrease in reflected light energy and an increase in scattered light energy, significantly enhancing the Raman signal [9].

Based on the CM mechanism, charge transfer efficiency occurs between R6G and the composite substrate [24]. Through XPS and UPS testing results, it can be concluded that the valence band maximum (VB) and conduction band minimum (CB) of the urchin-like Ta2O5 structure are positioned at -6.66 eV and -3.88 eV relative to the vacuum level, respectively. The HOMO and LUMO energy levels of R6G are -5.70 eV and -3.00 eV, respectively. Due to the HOMO-LUMO gap of R6G, which is approximately 1.90 eV, being significantly smaller than the energy of the incident 532 nm laser (2.33 eV), there is a clear energy level matching between R6G and Ta2O5, allowing for electron transfer from R6G's HOMO level to Ta2O5's conduction band and further to R6G's LUMO level. The Ta2O5 conduction band acts as a stepping stone for charge transfer efficiency and greatly facilitates the process. The energy level difference between the Fermi level of Ag nanoparticles (-4.84 eV) and the LUMO level of R6G and the conduction band minimum of Ta2O5, respectively, is approximately 1.84 eV and 0.96 eV, both significantly smaller than the energy of the incident laser (2.33 eV). Therefore, charge transfer efficiency also occurs between R6G and Ta2O5 and Ag nanoparticles(Fig. 8). This series of charge transfer efficiency effectively reduces the difficulty of electron transfer between R6G's HOMO and LUMO orbitals, increasing the scattering probability and enhancing the SERS effect of the substrate.

 figure: Fig. 8.

Fig. 8. The charge transfer efficiency between R6G, urchin-like Ta2O5 and Ag NPs.

Download Full Size | PDF

The urchin-like nanostructure of Ta2O5 also serves as a crucial foundation for achieving high sensitivity detection. It increases the substrate's surface area, providing more contact sites with R6G molecules. Moreover, the unique structure enables a tight integration between Ta2O5 and Ag NPs, facilitating charge transfer efficiency between the analyte molecules and the substrate, which contributes to enhanced Raman signals. Additionally, Ta2O5's excellent optical properties, combined with its special structure, allow for multiple reflections and refractions of light, enhancing the utilization efficiency of the laser and increasing scattering, ultimately boosting the Raman signal [40].

4. Conclusions

In this study, urchin-like Ta2O5 nanomaterials were successfully synthesized using hydrothermal reactions and combined with Ag NPs to create SERS substrates for detecting low concentration target molecules such as R6G, CV, and MB. The urchin-like structure of the Ta2O5 nanomaterials, with its unique structural characteristics and self-doping effect, effectively modulated the band structure, thereby improving the charge transfer efficiency and surface plasmon resonance effect. This provides a crucial foundation for achieving high sensitivity in SERS detection. Furthermore, the constructed SERS substrates demonstrated strong quantitative capability, stability under strong acid and alkali conditions, and good stability and recyclability, making them highly promising for applications in complex environments such as wastewater monitoring.

Funding

Natural Science Foundation of Shandong Province (ZR2020MA072).

Disclosures

The authors declare no conflicts of interest.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

References

1. K. Kneipp, H. Kneipp, I. Itzkan, et al., “Ultrasensitive chemical analysis by Raman spectroscopy,” Chem. Rev. 99(10), 2957–2976 (1999). [CrossRef]  

2. M. G. Albrecht and J. A. Creighton, “Anomalously intense Raman spectra of pyridine at a silver electrode,” J. Am. Chem. Soc. 99(15), 5215–5217 (1977). [CrossRef]  

3. L. Zha, X. H. Fang, Y. Han, et al., “Controlled fiber core mode and surface mode interaction for enhanced SERS performance,” Opt. Express 30(25), 44827–44836 (2022). [CrossRef]  

4. S. Nie and S. R. Emory, “Probing Single Molecules and Single Nanoparticles by Surface-Enhanced Raman Scattering,” Science 275(5303), 1102–1106 (1997). [CrossRef]  

5. B. Yang, S. Jin, S. Guo, et al., “Recent Development of SERS Technology: Semiconductor-Based Study,” ACS Omega 4(23), 20101–20108 (2019). [CrossRef]  

6. X. W. Xiu, L. P. Hou, J. Yu, et al., “Manipulating the surface-enhanced Raman spectroscopy (SERS) activity and plasmon-driven catalytic efficiency by the control of Ag NP/graphene layers under optical excitation,” Nanophotonics 10(5), 1529–1540 (2021). [CrossRef]  

7. C. Zhang, C. H. Lia, J. Yu, et al., “SERS activated platform with three-dimensional hot spots and tunable nanometer gap,” Sens. Actuators, B 258, 163–171 (2018). [CrossRef]  

8. H. K. Yu, Y. S. Peng, Y. Yang, et al., “Plasmon-enhanced light-matter interactions and applications,” npj Comput. Mater. 5(1), 45 (2019). [CrossRef]  

9. S. Y. Ding, J. Yi, J. F. Li, et al., “Nanostructure-based plasmon-enhanced Raman spectroscopy for surface analysis of materials,” Nat. Rev. Mater. 1(6), 16021 (2016). [CrossRef]  

10. J. R. Lombardi, R. L. Birke, T. Lu, et al., “Charge-transfer theory of surface enhanced Raman spectroscopy: Herzberg-Teller contributions,” J. Chem. Phys. 84(8), 4174–4180 (1986). [CrossRef]  

11. A. K. Samal, L. Polavarapu, S. Rodal-Cedeira, et al., “Size Tunable Au@Ag Core Shell Nanoparticles: Synthesis and Surface-Enhanced Raman Scattering Properties,” Langmuir 29(48), 15076–15082 (2013). [CrossRef]  

12. X. M. Qian, X. H. Peng, D. O. Ansari, et al., “In vivo tumor targeting and spectroscopic detection with surface-enhanced Raman nanoparticle tags,” Nat. Biotechnol. 26(1), 83–90 (2008). [CrossRef]  

13. X. T. Wang and L. Guo, “SERS Activity of Semiconductors: Crystalline and Amorphous Nanomaterials,” Angew. Chem., Int. Ed. 59(11), 4231–4239 (2020). [CrossRef]  

14. X. Ling, L. M. Xie, Y. Fang, et al., “Can Graphene be used as a Substrate for Raman Enhancement?” Nano Lett. 10(2), 553–561 (2010). [CrossRef]  

15. S. Z. Oo, S. Siitonen, V. Kontturi, et al., “Disposable gold coated pyramidal SERS sensor on the plastic platform,” Opt. Express 24(1), 724–731 (2016). [CrossRef]  

16. Y. T. Kim, J. Schilling, S. L. Schweizer, et al., “Morphology Dependence on Surface-Enhanced Raman Scattering Using Gold Nanorod Arrays Consisting of Agglomerated Nanoparticles,” Plasmonics 12(1), 203–208 (2017). [CrossRef]  

17. J. Krajczewski, S. Turczyniak-Surdacka, M. Dziubaltowska, et al., “Ordered zirconium dioxide nanotubes covered with an evaporated gold layer as reversible, chemically inert and very efficient substrates for surface-enhanced Raman scattering (SERS) measurement,” Spectrochim. Acta, Part A 275, 121183 (2022). [CrossRef]  

18. J. W. Luo, A. Franceschetti, and A. Zunger, “Carrier Multiplication in Semiconductor Nanocrystals: Theoretical Screening of Candidate Materials Based on Band-Structure Effects,” Nano Lett. 8(10), 3174–3181 (2008). [CrossRef]  

19. Y. Shan, Z. Zheng, J. Liu, et al., “Niobium pentoxide: A promising surface-enhanced Raman scattering active semiconductor substrate,” npj Comput. Mater. 3(1), 11 (2017). [CrossRef]  

20. L. L. Yang, Y. S. Peng, Y. Yang, et al., “Green and Sensitive Flexible Semiconductor SERS Substrates: Hydrogenated Black TiO2 Nanowires,” ACS Appl. Nano Mater. 1(9), 4516–4527 (2018). [CrossRef]  

21. J. R. Lombardi and R. L. Birke, “Theory of Surface-Enhanced Raman Scattering in Semiconductors,” J. Phys. Chem. C 118(20), 11120–11130 (2014). [CrossRef]  

22. Z. Zhang, Z. Li, L. Wang, et al., “Preparation and surface-enhanced Raman scattering properties of GO/Ag/Ta2O5 composite substrates,” Opt. Express 29(21), 34552–34564 (2021). [CrossRef]  

23. V. Gurylev, “A review on the development and advancement of Ta2O5 as a promising photocatalyst,” Materials Today Sustainability 18, 100131 (2022). [CrossRef]  

24. L. Yang, Y. Yang, J. R. Lombardi, et al., “Charge transfer enhancement in the surface -enhanced Raman scattering of Ta2O5 superstructures,” Appl. Surf. Sci. 520, 146325 (2020). [CrossRef]  

25. L. L. Yang, Y. S. Peng, Y. Yang, et al., “A Novel Ultra-Sensitive Semiconductor SERS Substrate Boosted by the Coupled Resonance Effect,” Adv. Sci. 6(12), 1900310 (2019). [CrossRef]  

26. L. D. Yuan, H. X. Deng, S. S. Li, et al., “Unified theory of direct or indirect band-gap nature of conventional semiconductors,” Phys. Rev. B 98(24), 245203 (2018). [CrossRef]  

27. L. Y. Wang, M. J. Liu, B. Chen, et al., “Graphene oxide/Ag nanoparticle/WS2 nanosheet heterostructures for surface-enhanced Raman spectroscopy,” Opt. Mater. Express 12(9), 3718–3730 (2022). [CrossRef]  

28. Z. Wang, J. Hou, C. Yang, et al., “Hierarchical metastable γ-TaON hollow structures for efficient visible-light water splitting,” Energy Environ. Sci. 6(7), 2134–2144 (2013). [CrossRef]  

29. T. Balankura, X. Qi, Y. Zhou, et al., “Predicting kinetic nanocrystal shapes through multi-scale theory and simulation: Polyvinylpyrrolidone-mediated growth of Ag nanocrystals,” J. Chem. Phys. 145(14), 144106 (2016). [CrossRef]  

30. X. J. Lu, S. J. Ding, T. Q. Lin, et al., “Ta2O5 Nanowires: a novel synthetic method and their solar energy utilization,” Dalton Trans. 41(2), 622–627 (2012). [CrossRef]  

31. J. Y. Duan, W. D. Shi, L. L. Xu, et al., “Hierarchical nanostructures of fluorinated and naked Ta2O5 single crystalline nanorods: hydrothermal preparation, formation mechanism and photocatalytic activity for H-2 production,” Chem. Commun. 48(58), 7301–7303 (2012). [CrossRef]  

32. X. Yu, W. Li, Z. Li, et al., “Defect engineered Ta2O5 nanorod: One-pot synthesis, visible-light driven hydrogen generation and mechanism,” Appl. Catal., B 217, 48–56 (2017). [CrossRef]  

33. X. Jiang, L. Xu, W. Ji, et al., “One plus one greater than Two: Ultrasensitive Surface-Enhanced Raman scattering by TiO2/ZnO heterojunctions based on Electron-Hole separation,” Appl. Surf. Sci. 584, 152609 (2022). [CrossRef]  

34. Q. H. Wu, A. Thissen, W. Jaegermann, et al., “Photoelectron spectroscopy study of oxygen vacancy on vanadium oxides surface,” Appl. Surf. Sci. 236(1-4), 473–478 (2004). [CrossRef]  

35. J. P. Wang, Z. Y. Wang, B. B. Huang, et al., “Oxygen Vacancy Induced Band-Gap Narrowing and Enhanced Visible Light Photocatalytic Activity of ZnO,” ACS Appl. Mater. Interfaces 4(8), 4024–4030 (2012). [CrossRef]  

36. F. C. Lei, Y. F. Sun, K. T. Liu, et al., “Oxygen Vacancies Confined in Ultrathin Indium Oxide Porous Sheets for Promoted Visible-Light Water Splitting,” J. Am. Chem. Soc. 136(19), 6826–6829 (2014). [CrossRef]  

37. Z. L. Shi, X. L. Hao, and C. X. Xu, “In situ synthesis of Ag nanoparticles-graphene oxide nanocomposites with strong SERS activity,” Mater. Res. Express 5(1), 015034 (2018). [CrossRef]  

38. X. W. Xiu, Y. Guo, C. H. Li, et al., “High-performance 3D flexible SERS substrate based on graphene oxide/silver nanoparticles/pyramid PMMA,” Opt. Mater. Express 8(4), 844–857 (2018). [CrossRef]  

39. C. Zhang, S. Z. Jiang, C. Yang, et al., “Gold@silver bimetal nanoparticles/pyramidal silicon 3D substrate with high reproducibility for high-performance SERS,” Sci. Rep. 6(1), 25243 (2016). [CrossRef]  

40. Y. D. Wang, N. Lu, W. T. Wang, et al., “Highly effective and reproducible surface-enhanced Raman scattering substrates based on Ag pyramidal arrays,” Nano Res. 6(3), 159–166 (2013). [CrossRef]  

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (8)

Fig. 1.
Fig. 1. Preparation process of Ag NPs/ Ta2O5 composite SERS substrate.
Fig. 2.
Fig. 2. (a-b) SEM image of Ag NPs, urchin-like Ta2O5. (c-d) SEM image of Ag NPs/Ta2O5 composite substrate. (e-f) TEM and HRTEM image of urchin-like Ta2O5. (g) electron diffraction pattern of urchin-like Ta2O5.
Fig. 3.
Fig. 3. (a-b) EDS image of urchin-like Ta2O5. (c) UV-Vis patterns of Ag NP, R6G and R6G on Ag NPs/Ta2O5. (d) XRD patterns of urchin-like Ta2O5. (e-f) UV-Vis patterns of urchin-like Ta2O5 showing wavelength and energy.
Fig. 4.
Fig. 4. (a-b) XPS spectra of Ta4f, O1s of urchin-like Ta2O5. (c) XPS survey spectrum of Ag NPs/ Ta2O5 composite substrate. (d-e) XPS spectra of Ta4f, O1s of purchased Ta2O5. (f) Valence-band XPS spectra of purchased Ta2O5 and urchin-like Ta2O5. (g-h) UPS spectra of urchin-like Ta2O5.
Fig. 5.
Fig. 5. (a) The Raman spectra of R6G on Ta sheet, purchased Ta2O5, urchin-like Ta2O5, Ag NPs and Ag NPs/Ta2O5 composite substrate using a laser power of 24 µW. (b) Raman spectra of R6G on the Ag NPs/Ta2O5 composite substrates from 10−7 M to 10−12 M using a laser power of 4.8 µW. (c) Ag NPs/Ta2O5 composite substrate's lowest detection limit for R6G. (d) The relationship between R6G concentration and the intensities of the Raman peaks at 611 cm-1 and 1362 cm-1. (e) Raman mapping image of R6G on the composite substrate. (f-g) Raman spectra of CV and MB on Ag NPs/Ta2O5 composite substrate at concentrations ranging from 10−7 M to 10−12 M.
Fig. 6.
Fig. 6. (a) The Raman spectra of 10−2 M R6G on SiO2 and 10−12 M R6G on Ag NPs/Ta2O5. (b-c) The results of the 30-day stability test of the composite substrate. (d) SEM image of Ag NPs/Ta2O5 substrate after HNO3 cleaning. (e) Comparison of Raman spectra before and after recycling treatment.
Fig. 7.
Fig. 7. Simulated electric field distribution of Ta2O5, Ag NPs, urchin-like Ta2O5 and Ag NPs/Ta2O5 composite substrates.
Fig. 8.
Fig. 8. The charge transfer efficiency between R6G, urchin-like Ta2O5 and Ag NPs.

Tables (1)

Tables Icon

Table 1. Sensitivity Comparison of Several Ta2O5-based or Ag NPs-based Composite SERS Substrates

Equations (5)

Equations on this page are rendered with MathJax. Learn more.

2 Ta + 14 HF 2 H 2 Ta F 7 + 5 H 2
2 H 2 Ta F 7 + 5 H 2 O T a 2 O 5 + 14 HF
H 2 O 2 + H 2 2 H 2 O
R S D = S I ¯ × 100 % = i = 1 n ( I i I ¯ ) 2 I
E F = I S E R S / N S E R S I S i O 2 / N S i O 2
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.