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

Preparation and surface-enhanced Raman scattering properties of GO/Ag/Ta2O5 composite substrates

Open Access Open Access

Abstract

The composite substrate composed of precious metal, semiconductor and graphene has not only high sensitivity and uniform Raman signal but also stable chemical properties, which is one of the important topics in the field of surface-enhanced Raman scattering (SERS). In this paper, a sandwich SERS substrate based on tantalum oxide (Ta2O5) is designed and fabricated. The substrate has high sensitivity, stable performance and high quantification capability. The composite substrate can achieve a high sensitivity Raman detection of crystal violet (CV) with a detection limit of 10−11 M and an enhancement factor of 1.5 × 109. This is the result of the synergistic effect of electromagnetic enhancement and chemical enhancement, in which the chemical enhancement is the cooperative charge transfer in the system composed of probe molecules, silver nanoparticles (AgNPs) and Ta2O5, and the electromagnetic enhancement comes from the strong local surface plasmon resonance between the adjacent AgNPs. After exposing the composite substrate to the air for one month, the Raman signal did not weaken, indicating that the performance of the composite substrate is stable. In addition, there is an excellent linear relationship between the intensity of Raman characteristic peak and the concentration of probe molecules, which proves that the composite substrate has high quantification capability. In practical application, the composite SERS substrate can be used to detect harmful malachite green quickly and sensitively and has a broad application prospect in the field of food safety and chemical analysis.

© 2021 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Raman spectrum is a kind of scattering spectrum which distinguishes the molecular structure of matter by analyzing the scattered light which is different from the frequency of incident light [1]. However, the scattered light intensity of the traditional Raman spectrum is very weak, which limits the development of the traditional Raman spectrum [2]. Surface-enhanced Raman scattering (SERS) is a technique that significantly enhances the Raman signals of molecules adsorbed on special substrates to identify chemical and biological molecules [3]. As a powerful and effective analytical technology, SERS has been widely used in food safety, medical diagnosis, environmental monitoring, chemical analysis and other fields [411]. At present, the generally accepted mechanisms of SERS enhancement include electromagnetic enhancement mechanism (EM) and chemical enhancement mechanism (CM), and EM plays a major role in the SERS effect [12]. For precious metal materials, when incident light of a certain frequency irradiates its surface, due to the coupling of free electrons on its surface and incident light, precious metal nanoparticles will produce a strong local surface plasmon resonance (LSPR), resulting in SERS based on EM [13,14]. The precious metals (such as Au and Ag) SERS substrate based on EM have significant SERS effect, and the enhancement factor is as high as 1014, which can realize single molecule detection [15]. For non-precious metal materials, there are three main situations: (1) The CM of most semiconductor SERS substrates mainly comes from the enhancement of charge transfer, which is closely related to the inherent properties of the adsorbed molecules and the surface properties of non-noble metal substrates. When the highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO) match with the valence band and conduction band energy levels of the substrate, there is an effective charge shift, which changes the polarizability tensor and electron density distribution of the molecule, thus the SERS effect is observed [16]. (2) For some metal-like materials such as MoO2, transition metal Cu, etc., the free electron density of the surface can be adjusted by changing its microstructure, thus generating LSPR in the visible region to improve the SERS performance of the substrate [17]. (3) Mie scattering occurs when the size of the nanostructure or cavity structure on the substrate is equal to the wavelength of the incident light, which can also improve the SERS performance of the substrate [18]. Common non-precious metal materials such as CdTe, TiO2, ZnO, Cu2O, Ta2O5 have been developed into SERS active substrate materials [1922]. They enhance the performance of SERS by changing the shape and structure or by doping with self-doping and hetero-doping. Among them, Ta2O5 nanoparticles have the advantages of high biocompatibility, large specific surface area and high chemical stability, so they are one of the semiconductor substrates with the highest enhancement factor [2326]. In addition, graphene, as an outstanding representative of two-dimensional materials, has been widely used in the study of SERS substrates because of its strong oxidation resistance, strong light transmittance, high electron mobility, excellent heat conduction performance, high toughness and strong chemical stability [2731].

However, with the continuous development of SERS technology in practical application, the shortcomings of traditional substrates are increasingly exposed because of the single preparation material, such as poor spectral stability and weak repeatability of precious metal substrates [32], low sensitivity and weak enhancement effect of non-precious metal substrates [33]. Therefore, to solve the limitations of single precious metal substrate and non-precious metal substrate and improve the substrate performance to meet the practical application is one of the important topics in the future SERS field [34]. Compared with the substrate of single nanomaterial, the composite substrate composed of two or more materials not only has high sensitivity, but also has uniform Raman signal, stable chemical properties and high biocompatibility [35,36]. In the metal-semiconductor structure, the precious metal shows a strong surface plasmon resonance effect in the visible region, which greatly enhances the SERS effect. Among the common precious metal nanostructures, silver nanoparticles (AgNPs) has excellent SERS properties and controllable morphology, so it has been widely used in the preparation of metal-semiconductor composite substrates [37,38]. At the same time, semiconductor materials, as the bottom or covering layer of metal nanostructures, can increase the specific surface area, which is conducive to the adsorption of the measured molecules, thus enhancing the Raman signal of the tested molecules. In addition, it can avoid the oxidation of precious metal nanostructures directly exposed to air, thus enhancing the stability of the substrate. Based on this theory, the composite substrates of graphene and precious metal materials show excellent SERS properties [3940]. Compared with graphene, graphene oxide (GO) contains a large number of oxygen-containing functional groups and has higher surface activity, and can combine well with precious metals and semiconductors [4144]. The multilayer GO not only has the same charge transfer mechanism of π-π stacking as graphene, but also the oxygen-containing groups can adsorb dye molecules, thus enhancing the SERS effect. Therefore, the SERS performance of GO-based composite substrate is better than that of graphene [4547]. To sum up, GO composite substrate has become a hot spot in the research and preparation of SERS substrate because of its excellent performance and potential application value.

In addition, it has been reported that the sensitivity and quantitative capability of SERS are mutually exclusive [48], because the contribution of analyte molecules at different positions to SERS is very different, which inevitably leads to poor quantitative capability of [49]. In order to solve this problem, we improve the quantitative capability of the Ag/Ta2O5 composite substrate by reducing the size of AgNPs and increasing the number of hot spots per unit volume. Ta2O5 and AgNPs with smaller size were prepared by hydrothermal reaction. Combined with the advantages of precious metal materials and semiconductor materials, Ag/Ta2O5 composite substrate with sandwich structure were designed. The Ta2O5 placed at the bottom can attach uniform AgNPs to establish high density active hot spots. In addition, the hydrothermal reaction makes Ta2O5 surface bound with a large number of hydroxyl groups, which makes it have stronger electrostatic attraction to basic dyes molecule, thus improving the Raman signal of the measured molecules. We use different concentrations of rhodamine 6G (R6G) to detect the sensitivity and quantification capability of different batches of Ag/Ta2O5 composite substrates, and the mapping diagram is used to reflect the consistency of the substrate performance. Through experiments, we believe that the SERS enhancement of the composite substrate is the synergistic effect of electromagnetic enhancement and chemical enhancement, in which the chemical enhancement is the cooperative charge transfer in the system composed of probe molecules and Ag/Ta2O5 substrate, and the electromagnetic enhancement comes from the strong local surface plasmon resonance between the adjacent AgNPs. We also discuss the stability of the composite substrate with or without GO. The results show that GO can be used as an inert film to prevent AgNPs from being oxidized, which is beneficial to improve the stability of the composite substrate. In practical application, the composite substrate is still sensitive to malachite green (MG) solution prepared with lake water. Finally, we use the finite-difference time-domain method (FDTD) to simulate the experiment, and the results are consistent with our experiments. Our work shows that GO/Ag/Ta2O5 composite SERS substrate has a broad application prospect in the field of food safety and chemical analysis.

2. Experimental details

2.1 Materials

Tantalum oxide (Ta2O5, 99.9% metals basis), hydrogen peroxide solution (H2O2, AR, 30 wt. % in H2O), silver nitrate (AgNO3, AR, 99.8%), hydrofluoric acid (HF, AR, ≥40%), polyvinylpyrrolidone (PVP, average Mw 10000) were purchased from Aladdin Co., Ltd. (Shanghai, China). Ethanol absolute (C2H6O, AR, ≥99.7%), acetone (C3H6O, AR, ≥99.5%), ethylene glycol (C2H6O2, AR, ≥99.5%) and ammonium hydroxide aqueous solution (NH3·H2O, AR, 25∼28%) were bought from Sinopharm Chemical Reagent Co., Ltd. GO dispersion was prepared by Hummers method [50]. All the water used in the experiment is deionized water.

2.2 Synthesis of Ta2O5 nanoparticles and AgNPs

Ta2O5 nanoparticles were prepared by hydrothermal synthesis in two steps [23]. The first step is to dissolve 2.0 g Ta2O5 powder in 20 ml HF and then keep it at 110 °C for 12 hours. Subsequently, NH3·H2O is dropped into the mixed solution to form a large number of white flocculent precipitates. Then the precipitates were cleaned and dried to obtain Ta2O5 precursors. The Ta2O5 precursor is the Ta2O5 powder purified from the purchased Ta2O5 powder. In the second step, the 0.04 g Ta2O5 precursor was fully dissolved in the mixed solution of 50 ml H2O2 and 10 ml NH3·H2O, then the mixed solution was transferred to a hydrothermal reactor and a silicon wafer was added at the same time and kept at 240 °C for 36 hours. Finally, the white powdered Ta2O5 nanoparticles were obtained by cleaning and drying.

AgNPs were prepared by hydrothermal synthesis method [51]. First, 0.5 g PVP and 0.05 g AgNO3 were dissolved in 20 ml ethylene glycol solution. Then heat the mixed solution to 130 °C and hold for 1.5 hours. When the mixed solution turns bright yellow, stop heating and add 45 ml of acetone to it. Finally, AgNPs were obtained by centrifugation and cleaning.

2.3 Preparation of the composite substrate

The preparation process of GO/Ag/Ta2O5 composite SERS substrate is shown in the Fig. 1. Firstly, 0.002 g Ta2O5 nanoparticles were dissolved in 20 ml deionized water to form Ta2O5 solution. 30 μl Ta2O5 solution was dripped on the hydrophilic quartz sheet, then heated and solidified to form the bottom layer of Ta2O5. A layer of AgNPs was deposited on the Ta2O5 substrate by impregnation method. Finally, the GO/Ag/Ta2O5 composite substrate was obtained by spin-coating GO dispersion on the Ag/Ta2O5 composite substrate.

 figure: Fig. 1.

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

Download Full Size | PDF

2.4 Experimental instrument

Field emission scanning electron microscopy (SEM, ZEISS Gemini Sigma 500) is used to characterize the morphology and structure of composite substrates. Ultraviolet and visible spectrophotometer (UV-Vis, Alpha-1500) is used to measure the absorption spectra of AgNPs solution. Raman detection were performed by using a Raman spectrometer (Horiba HR Evolution 800) at a laser wavelength of 532 nm. The laser power was set as 0.48 mW with an integration time of 4 s and the diffraction grating is 600 gr/mm. When the laser was irradiated on the substrate, the laser power was 0.48 mW and diameter of the spot was about 1µm. Raman spectra were randomly collected on 15 positions, the average values were calculated. Throughout the SERS experiment, the sample was observed with a microscope equipped with a 50-fold objective.

3. Results and discussion

3.1 Characterization of the composite substrate

The morphology and structure of GO/Ag/Ta2O5 composite substrate were characterized by scanning electron microscope (SEM). As shown in the Fig. 2(a), it can be observed that the prepared Ta2O5 nanoparticles are uniform in size and spherical in shape. We use image J software to analyze the particle diameter of SEM photos. We calculate the particle diameter in the range of 1.0 × 0.8 μm, and there are 204 Ta2O5 nanoparticles in this area. The results (Fig. 2(h)) show that the average diameter of Ta2O5 nanoparticles is 28.4 nm, and the Ta2O5 nanoparticles with particle diameter distribution between 25-30 nm account for 41.2% of the total. These uniform Ta2O5 nanoparticles can make Ag/Ta2O5 composite substrate have high density and uniform SERS hot spots. As shown in Fig. 2(b), the morphology and structure of AgNPs were characterized by SEM. The prepared AgNPs are uniform in size and spherical in shape, which is beneficial to improve the plasma coupling efficiency in the later stage [52]. Similarly, we calculate the particle diameter in the range of 3.0 × 3.0 μm, and there are 600 AgNPs in this area. The results (Fig. 2(i)) show that the average particle size of AgNPs is 72.9 nm. As shown in Fig. 2(c), we can see that the AgNPs in the Ag/Ta2O5 composite substrate is closely arranged, and the Ta2O5 nanoparticles at the bottom can be seen from the gap of the Ag. In the white circle marked (Fig. 2(d)), we can clearly see the existence of gauze-like GO film. In Fig. 2(e), the GO/Ag/Ta2O5 composite substrate covered with GO film and the Ag/Ta2O5 composite substrate without GO film can be clearly distinguished. The element distribution of GO/Ag/Ta2O5 composite substrate was analyzed by energy dispersion spectrometer (EDS). As shown in Fig. 2(f), the standard diffraction peaks of Ta, Ag, C, Si and O further confirm the composite structure of Ta2O5, Ag and GO. As shown in Fig. 2(g), when the local composition of the sample is measured by EDS element mapping, we can clearly see the uniform distribution of Ta, Ag, C and O on the substrate, which proves that GO, Ag and Ta2O5 are uniformly distributed on the quartz substrate. Figure 2(j) shows the absorption peak of the visible part is about 455 nm, so the SERS signal with high signal-to-noise ratio can be produced under the incident laser with 532 nm wavelength. The full width at half maxima (FWHM) is only 100 nm, which indicates that the prepared nanoparticles are uniform. Figure 2(k) shows the Raman spectrum of GO on the GO/Ag/Ta2O5 composite substrate. The D peak at 1353 cm-1 is half of the G peak at 1583 cm-1 intensity and has a large bandwidth, indicating that GO has obvious structural disorder. There is no obvious 2D peak at 2800 cm-1, on the contrary, a defect activation peak called Ding can be seen at 2940 cm-1, which also indicates that the structure of GO is disordered [53].

 figure: Fig. 2.

Fig. 2. (a-d) SEM image of Ta2O5, AgNPs, Ag/Ta2O5 and GO/Ag/Ta2O5 composite substrates. (e) SEM image of Ag/Ta2O5 and GO/Ag/Ta2O5 composite substrates. (f) The energy dispersion spectrometer of the GO/Ag/Ta2O5 composite substrates. (g) EDS elemental maps from the GO/Ag/Ta2O5 composite substrates. (h,i) The size distribution of Ta2O5 and AgNPs. (j) The UV-Vis absorption spectrum of AgNPs. (k) The Raman spectrum of GO.

Download Full Size | PDF

The chemical states of Ag, Ta and O in purchased Ta2O5 and Ag/Ta2O5 composite substrate were analyzed by X-ray photoelectron spectroscopy (XPS). The standard value 284.8 eV of C1s is used to correct the charge of XPS energy spectrum. Figure (a,b), as a control group, shows the XPS energy spectrum of the purchased Ta2O5. As shown above, the Ta4f spectra of purchased Ta2O5 (Fig. 3(a)) and Ag/Ta2O5 (Fig. 3(d)) composite substrates can be divided into two peaks, belonging to the characteristic peaks Ta5+ 4f7/2 and Ta5+ 4f5/2. Compared with purchased Ta2O5, the Ta4f peak of Ag/Ta2O5 composite substrate moves slightly to higher binding energy, from 27.8 eV and 26 eV to 28 eV and 26.2 eV, this may be the charge transfer between Ta2O5 and AgNPs. In addition, compared with purchased Ta2O5 (Fig. 3(b)), the proportion of Ta-OH in Ag/Ta2O5 (Fig. 3(e)) composite substrate is much larger. This shows that a large number of hydroxyl groups (-OH) are bound to the surface of Ta2O5 nanostructures after hydrothermal reaction. The -OH groups (function as Bronsted acid sites) on the surface of Ta2O5 nanostructures make them more electrostatically attractive to the basic dye R6G, CV and MG molecules [24]. Therefore, it directly affects the direction and efficiency of charge transfer between the composite substrate and dye molecules, thus improving the enhancement effect of SERS. As shown in Fig. 3(c,f), the signals clearly observed at 367.8 eV and 373.8 eV are Ag 3d5/2 and Ag 3d3/2, indicating the presence of AgNPs.

 figure: Fig. 3.

Fig. 3. (a,b) Ta4f and O1s detailed XPS spectra of purchased Ta2O5 (c) Ag3d detailed XPS spectra of Ag/Ta2O5 composite substrate. (d,e) Ta4f and O1s detailed XPS spectra of Ag/Ta2O5 composite substrate.(f) Survey XPS spectra of Ag/Ta2O5 composite substrate.

Download Full Size | PDF

3.2 SERS performance of the composite substrates

We select R6G dye molecules for limit testing on SiO2, Ta2O5, AgNPs and Ag/Ta2O5 composite substrates. As shown in Fig. 4(a), the characteristic peak of R6G on AgNPs and Ag/Ta2O5 composite substrates is much stronger than that of Ta2O5, which can be attributed to the strong LSPR produced by dense and uniform AgNPs. The comparison between Ta2O5 and SiO2 shows that Ta2O5 has better SERS activity. In addition, the characteristic peak of R6G on Ag/Ta2O5 composite substrate is slightly stronger than that of AgNPs. This is attributed to the charge transfer between R6G and Ag/Ta2O5, which increases the SERS effect. Figure 4(b) shows the Raman spectra of different concentrations of R6G on Ag/Ta2O5 composite substrates prepared under the same experimental conditions. When the concentration of R6G is 10−12 M, we can still clearly distinguish the main characteristic peaks of R6G molecules, such as 613, 1363 and 1511 cm-1. As shown in Fig. 4(c), we analyze the relationship between the intensity of the R6G peak and the molecular concentration, and the error strip represents the standard deviation of 15 spectra. The linear fitting was carried out with the concentration as the variable and the peak intensity as the dependent variable. The results showed that the coefficient of determination (R2) at 613 cm-1 and 1364 cm-1 were 0.989 and 0.969, respectively. Based on the fitting line, we can quantify the R6G dye molecules to a certain extent based on Raman spectra. Figure 4(d) shows the intensity distribution of the characteristic peaks of R6G molecules obtained on Ag/Ta2O5 composite substrates at 613, 1364 and 1511 cm-1. According to the formula of relative standard deviation (RSD):

$$RSD = \frac{S}{\bar{x}} \times 100\%= \frac{{\sqrt {\frac{{\mathop \sum \nolimits_{i = 1}^n {{\left( {{x_i} - \bar{x}} \right)}^2}}}{{n - 1}}} }}{{\bar{x}}} \times 100\%.$$

In the formula, S is the standard deviation, $\bar{x}$ represents the corresponding average value, i represents each measured value, and n represents the number of tests. Through the calculation, the RSD is 4.47%, 4.67% and 7.23% respectively, demonstrating consistent performance between composite substrates. Figure 4(e) shows the Raman mapping of the characteristic peak at 613 cm-1 obtained in the range of 20 × 20 μm on the Ag/Ta2O5 composite substrate. The relatively uniform color of the spectrum shows that the intensity of the Raman characteristic peak is uniform in most ranges, which is attributed to the fact that the bottom Ta2O5 can attach uniform AgNPs, so that the composite substrate has a high uniformity. Figure 5(f) shows the charge transfer between R6G, Ta2O5 and AgNPs. There is a charge transfer effect between R6G and Ta2O5. The HOMO and the LUMO energy levels of R6G are -5.70 eV and -3.40 eV [54], respectively. The valence band (VB) and conduction band (CB) of Ta2O5 nanoparticles are -7.91 eV and -3.78 eV, respectively. Because the band gap between the HOMO level of R6G molecule and the conduction band of Ta2O5 nanoparticles is about 1.92 eV, which is obviously smaller than the incident laser energy (2.33 eV). Therefore, electrons may be transferred from the HOMO level of R6G to the Ta2O5 conduction band. Secondly, there is charge transfer between R6G and Ta2O5 and AgNPs. This is due to the fact that the energy level difference between the Fermi level (-4.84 eV) of AgNPs and the LUMO level of R6G molecules is about 0.86 eV [38], and the energy level difference between AgNPs and the conduction band of Ta2O5 is 0.38 eV, which is obviously smaller than the incident laser energy (2.33eV). This series of charge transfer effectively increases the polarizability of the molecule, which in turn enhances the SERS effect of the substrate. Therefore, there is charge transfer in the system formed by R6G molecules and Ag/Ta2O5 composites, which plays an important role in the enhancement of SERS signals of R6G molecules.

 figure: Fig. 4.

Fig. 4. (a) The Raman spectra of different concentrations of R6G on SiO2, Ta2O5, AgNPs and Ag/Ta2O5 composite substrates. (b) Raman spectra of R6G on the Ag/Ta2O5 composite substrates from 10−7 M to 10−12 M. (c) Log−log plot of average intensity of SERS signals at 613 and 1363 cm−1 versus the concentration of R6G. (d) The intensity distributions of the 613, 1364 and 1511 cm−1 peaks for R6G. (e) Corresponding Raman mapping of the 613 cm−1 characteristic peak of R6G (10−7 M). (f) The charge transfer between R6G, Ta2O5 and AgNPs.

Download Full Size | PDF

 figure: Fig. 5.

Fig. 5. (a) The Raman spectra of different concentrations of CV on SiO2, Ag/Ta2O5 and GO/Ag/Ta2O5 composite substrates. (b) Raman spectra of CV on the GO/Ag/Ta2O5 composite substrates from 10−6 M to 10−11 M. (c) Log−log plot of average intensity of SERS signals at 915 and 1621 cm−1 versus the concentration of CV. (d) Raman spectra of CV (10−7 M) on GO/Ag/Ta2O5 composite substrates at different time. (e) Raman spectra of CV (915 cm-1) on GO/Ag/Ta2O5 composite substrate and Ag/Ta2O5 composite substrate. (f) Corresponding Raman mapping of the 915 cm−1 characteristic peak of CV (10−6 M).

Download Full Size | PDF

As a triarylmethane dye, Crystal Violet (CV) can be used as a disinfectant for human skin. However, this dye has carcinogenicity and is classified as a stubborn molecule because it is not easy to be metabolized by microorganisms and can persist in a variety of environments [55]. Therefore, it is necessary to detect CV accurately and quickly. As shown in Fig. 5(a), we used CV to test the limit of SiO2, Ag/Ta2O5 and GO/Ag/Ta2O5 composite substrate. When the concentration of CV is 10−11 M, the main characteristic peaks of CV can still be clearly distinguished. The intensity of the characteristic peak of CV on GO/Ag/Ta2O5 composite substrate is slightly stronger than that on Ag/Ta2O5 composite substrate, which can be attributed to the chemical enhancement caused by π-π stacking of GO films and charge transfer between oxygen-containing functional groups and CV molecules. Figure 5(b) shows the Raman spectra of different concentrations of CV on GO/Ag/Ta2O5 composite substrates prepared under the same experimental conditions. As shown in Fig. 4(c), we have analyzed the relationship between the intensity of the characteristic peak and the molecular concentration. The linear fitting was carried out with the concentration as the variable and the peak intensity as the dependent variable, and the error strip represents the standard deviation of 15 spectra. The results showed that the R2 at 915 and 1621 cm-1 were 0.995 and 0.990, respectively. Based on the fitting line, we can quantify the CV dye molecules to a certain extent based on Raman spectra. In order to verify the effect of GO on the stability of composite substrates, Ag/Ta2O5 and GO/Ag/Ta2O5 composite substrates were exposed to air and tested every 7 days. As shown in Fig. 5(d), it can be clearly observed that the spectral peak intensity of CV on the GO/Ag/Ta2O5 composite substrate attenuates slowly with the passage of time. As shown in Fig. 5(e), we analyze the relationship between the intensity of the characteristic peak of Ag/Ta2O5 and the GO/Ag/Ta2O5 composite substrate at 915 cm-1 and the time, and the error strip represents the standard deviation of 15 spectra. The signal on the Ag/Ta2O5 substrate decreased by 36.26%, while the signal on the GO/Ag/Ta2O5 composite substrate decreased by 20.49%. This phenomenon can be attributed to the protective effect of GO. GO film can be used as a barrier for atomic diffusion, which can effectively prevent the penetration of oxygen and protect AgNPs from oxidation, so that the GO/Ag/Ta2O5 composite substrate has good stability. Figure 5(f) shows the Raman mapping of the 915 cm-1 characteristic peak obtained on the GO/Ag/Ta2O5 composite substrate. The relatively uniform color of the spectrum shows that the intensity of the Raman characteristic peak is uniform in most ranges, indicating that the introduction of GO will not affect the uniformity of the composite substrate.

The SERS performance of the composite substrate can be evaluated intuitively by calculating the enhancement factor (EF). The EF for CV molecules is calculated by the standard equation [56]:

$$EF = \frac{{{I_{SERS}}/{N_{SERS}}}}{{{I_{RS}}/{N_{RS}}}}$$

In the formula, ISERS and IRS are the intensity of the SERS peak and the normal Raman intensity respectively, and NRS and NSERS are the number of CV molecules per unit volume of SERS and normal Raman scattering. The peak strength of 10−2 M CV at 915 cm-1 on SiO2 is 202.15 and the peak intensity of 10−11 M CV on GO/Ag/Ta2O5 composite substrate is 304.61. As a result, the average EF of GO/Ag/Ta2O5 composite substrate is 1.5 × 109, which is higher than that of several Ta2O5-based or GO/Ag-based composite SERS substrates that have been reported, as shown in Table 1.

Tables Icon

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

Malachite green (MG) is a common dye and fungicide, which is widely used in aquaculture because of its low cost. However, it is also a carcinogen, which can stay in fish for a long time and may cause human teratogenicity, carcinogenicity and mutagenicity through the food chain [57]. Therefore, it is necessary to detect MG accurately and quickly in aquatic products. In order to study the possibility of the application of GO/Ag/Ta2O5 composite substrate in practical samples, we carried out SERS detection of MG in deionized water and lake water. We dripped 0.2 μL of lake water on the quartz substrate and tested it after natural drying. As shown in Fig. 6(a), it can be seen that there is only one Raman characteristic peak (482 cm-1) from the quartz substrate in the Raman spectrum of the lake water. Similarly, the Raman spectra of 10−2 M concentration MG prepared with lake water also have Raman characteristic peaks from quartz substrate. Therefore, it can be proved that there is no obvious Raman characteristic peak in the lake water. We prepared MG solutions with different concentrations from lake water to test the practical application effect of GO/Ag/Ta2O5 composite substrates. In Fig. 6(b), we can easily detect the characteristic Raman peaks of MG prepared by lake water in the concentration range of 10−5 M to 10−10 M. In Fig. 6(c), we linearly fit the intensity of the MG peak at 1618 and 1368 cm-1 with the molecular concentration, and the error strip represents the standard deviation of 15 spectra. The results show that the R2 at 1618 and 1368 cm-1 is about 0.995 and 0.985. On the basis of the fitting line, the MG dye molecules in lake water can be quantified to a certain extent according to Raman spectra. Therefore, the above results show that the proposed GO/Ag/Ta2O5 composite substrate can meet the practical application, and show the great application potential of the composite substrate in the field of food safety and chemical detection.

 figure: Fig. 6.

Fig. 6. (a) The Raman spectra of the measured lake water without MG on SiO2, the Raman spectra of 10−2 M MG prepared with lake water on SiO2, and the Raman spectra of 10−10 M MG droplets prepared with lake water on the composite substrate. (b) Raman spectra of MG on the GO/Ag/Ta2O5 composite substrates from 10−5 M to 10−10 M. (c) Log−log plot of average intensity of SERS signals at 1618 and 1368 cm−1 versus the concentration of MG.

Download Full Size | PDF

In order to better understand the electric field enhancement mechanism of GO/Ag/Ta2O5 composite substrates, we use the finite-difference time-domain method (FDTD) software to analyze the local electric field distribution of these structures. According to the SEM image and the actual experimental conditions, we set the wavelength of incident light to 532 nm, the diameter of AgNPs to 72.9 nm the diameter of Ta2O5 particles to 28.4nm, and the thickness of GO to 1 nm to simulate. By comparing the Fig. 7(a,b), it is obvious that due to the strong LSPR between AgNPs, the local electric field between AgNPs is very strong. It can be seen from Fig. 7(c) that the local electric field strength between AgNPs and Ta2O5 is much higher than that in Fig. 7(a), which can be attributed to the chemical enhancement caused by the charge transfer between AgNPs and Ta2O5. This is due to the fact that the energy difference between the Fermi level of AgNPs and the conduction band of Ta2O5 is 0.38eV, which is obviously smaller than the incident laser energy (2.33eV), so the charge transfer occurs. As shown in Fig. 7(d), when GO film is introduced, there is no new hot spot between GO film and AgNPs. This is due to the fact that the GO monolayer we set up does not have π-π stacking and there is no charge transfer between the oxygen-containing functional groups and the probe molecules, so it does not cause any SERS enhancement. Through FDTD, we verified that AgNPs and Ta2O5 can also have charge transfer in the absence of probe molecules, which is consistent with our experimental results.

 figure: Fig. 7.

Fig. 7. (a)-(d) Simulated electric field distribution of Ta2O5, AgNPs, Ag/Ta2O5 and GO/Ag/Ta2O5 composite substrates.

Download Full Size | PDF

4. Conclusions

In this paper, we have prepared a GO/Ag/Ta2O5 composite substrate with high sensitivity, high quantification capability and stable properties. We use R6G and CV molecules to carry out Raman detection of the composite substrate under different conditions. It is proved that the reason for the high sensitivity is the synergistic effect of chemical enhancement and electromagnetic enhancement, and the reason for the stability comes from the protection of GO. At the same time, a large number of hydroxyl groups on the surface of Ta2O5 nanoparticles and oxygen-containing functional groups on the surface of GO enhance the molecular selectivity of the GO/Ag/Ta2O5 composite substrate. We also prove that the introduction of GO not only improves the stability of the composite substrate, but also keeps the performance of the substrate consistent. Based on the practical application, we verify the feasibility of detecting MG content in lake water with GO/Ag/Ta2O5 composite substrate, which provides a new choice for water quality and food detection.

Funding

Natural Science Foundation of Shandong Province (ZR201910280104).

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. F. W. King, R. P. Van Duyne, and G. C. Schatz, “Theory of Raman scattering by molecules adsorbed on electrode surfaces,” J. Chem. Phys. 69(10), 4472–4481 (1978). [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. B. Sharma, R. R. Frontiera, A. Henry, E. Ringe, and R. P. V. Duyne, “SERS: materials, applications, and the future,” Mater. Today 15(1-2), 16–25 (2012). [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. Y. Guo, J. Yu, C. H. Li, Z. Li, J. Pan, A. H. Liu, B. Y. Man, T. F. Wu, and X. W. Xiu, “SERS substrate based on the flexible hybrid of polydimethylsiloxane and silver colloid decorated with silver nanoparticles,” Opt. Express 26(17), 21784–21796 (2018). [CrossRef]  

6. K. Kneipp, H. Kneipp, I. Itzkan, R. R. Dasari, and M. S. Feld, “Surface-enhanced Raman scattering: a new tool for biomedical spectroscopy,” Curr. Sci. 77(7), 915–924 (1999).

7. G. Braun, S. J. Lee, M. Dante, T. Q. Nguyen, M. Moskovits, and N. Reich, “Surface-enhanced Raman spectroscopy for DNA detection by nanoparticle assembly onto smooth metal films,” J. Am. Chem. Soc. 129(20), 6378–6379 (2007). [CrossRef]  

8. L. P. Hou, M. R. Shao, Z. Li, X. F. Zhao, and Z. Li, “Elevating the density and intensity of hot spots by repeated annealing for high-efficiency SERS,” Opt. Express 28(20), 29357–29367 (2020). [CrossRef]  

9. X. F. Zhao, C. D. Liu, J. Yu, Z. Li, L. Liu, C. Li, S. Xu, W. Li, B. Man, and C. Zhang, “Hydrophobic multiscale cavities for high-performance and self-cleaning surface-enhanced Raman spectroscopy (SERS) sensing,” Nanophotonics 9(16), 4761–4773 (2020). [CrossRef]  

10. F. Pozzi and M. Leona, “Surface-enhanced Raman spectroscopy in art and archaeology,” J. Raman Spectrosc. 47(1), 67–77 (2016). [CrossRef]  

11. C. H. Li, S. C. Xu, J. Yu, Z. Li, W. F. Li, J. H. Wang, A. H. Liu, B. Y. Man, S. K. Yang, and C. Zhang, “Local hot charge density regulation: Vibration-free pyroelectric nanogenerator for effectively enhancing catalysis and in-situ surface enhanced Raman scattering monitoring,” Nano Energy 81, 105585 (2021). [CrossRef]  

12. X. W. Xiu, L. P. Hou, J. Yu, S. Z. Jiang, C. H. Li, X. F. Zhao, Q. Q. Peng, S. Qiu, C. Zhang, B. Y. Man, and Z. Li, “Manipulating the surface-enhanced Raman spectroscopy (SERS) activity and plasmon-driven catalytic efficiency by the control of Ag NP/graphene layers under optical excitation,” J. Nanophotonics 10(5), 1529–1540 (2021). [CrossRef]  

13. C. Zhang, C. H. Li, J. Yu, S. Z. Jiang, S. C. Xu, C. Yang, Y. J. Liu, X. G. Gao, A. H. Liu, and B. Y. Man, “SERS activated platform with three-dimensional hot spots and tunable nanometer gap,” Sensors and Actuators B: Chemical 258, 163–171 (2018). [CrossRef]  

14. X. Xiu, Y. Guo, C. Li, Z. Li, D. Li, C. Zang, S. Jiang, A. Liu, B. Man, and C. Zhang, “High-performance 3D flexible SERS substrate based on graphene oxide/silver nanoparticles/pyramid PMMA,” Opt. Mater. Express 8(4), 844–857 (2018). [CrossRef]  

15. K. Kneipp, Y. Wang, H. Kneipp, L. T. Perelman, I. Itzkan, R. R. Dasari, and M. S. Feld, “Single molecule detection using surface-enhanced Raman scattering (SERS),” Phys. Rev. Lett. 78(9), 1667–1670 (1997). [CrossRef]  

16. X. T. Wang and L. Guo, “SERS activity of semiconductors: crystalline and amorphous nanomaterials,” Angew. Chem. Int. Ed. 59(11), 4231–4239 (2020). [CrossRef]  

17. Q. Q. Zhang, X. S. Li, Q. Ma, Q. Zhang, H. Bai, W. C. Yi, J. Y. Liu, J. Han, and G. C. Xi, “A metallic molybdenum dioxide with high stability for surface enhanced Raman spectroscopy,” Nat. Commun. 8(1), 14903 (2017). [CrossRef]  

18. L. Shi, T. U. Tuzer, R. Fenollosa, and F. Meseguer, “A new dielectric metamaterial building block with a strong magnetic response in the sub-1.5-micrometer region: silicon colloid nanocavities,” Adv. Mater. 24(44), 5934–5938 (2012). [CrossRef]  

19. Y. Wang, J. Zhang, H. Jia, M. Li, J. Zeng, B. Yang, B. Zhao, W. Xu, and J.R. Lombardi, “Mercaptopyridine surface-functionalized CdTe quantum dots with enhanced raman scattering properties,” J. Phys. Chem. C 112(4), 996–1000 (2008). [CrossRef]  

20. X. Xue, W. Ji, Z. Mao, H. J. Mao, Y. Wang, X. Wang, W. D. Ruan, B. Zhao, and J. R. Lombardi, “Raman investigation of nanosized TiO2: effect of crystallite size and quantum confinement,” J. Phys. Chem. C 116(15), 8792–8797 (2012). [CrossRef]  

21. Q. Liu, L. Jiang, and L. Guo, “Precursor-directed self-assembly of porous ZnO nanosheets as high-performance surface-enhanced Raman scattering substrate,” Small 10(1), 48–51 (2014). [CrossRef]  

22. J. Lin, Y. Shang, X. X. Li, J. Yu, X. T. Wang, and L. Guo, “Ultrasensitive SERS detection by defect engineering on single Cu2O superstructure particle,” Adv. Mater. 29(5), 1604797 (2016). [CrossRef]  

23. L. L. Yang, Y. Yang, J. R. Lombardi, Y. Peng, and Z. G. Huang, “Charge transfer enhancement in the surface-enhanced Raman scattering of Ta2O5 superstructures,” Appl. Surf. Sci. 520, 146325 (2020). [CrossRef]  

24. L. L. Yang, Y. S. Peng, Y. Yang, J. J. Liu, H. L. Huang, B. H. Yu, J. M. Zhao, Y. L. Lu, Z. G. Huang, Z. Y. Li, and J. R. Lombardi, “A novel ultra-sensitive semiconductor SERS substrate boosted by the coupled resonance effect,” Adv. Sci. 6(12), 1900310 (2019). [CrossRef]  

25. J. X. Li, W. L. Dai, J. Q. Yan, G. J. Wu, L. D. Li, and N. J. Guan, “Hydrothermal synthesis and photocatalytic properties of tantalum pentoxide nanorods,” Chinese Journal of Catalysis 36(3), 432–438 (2015). [CrossRef]  

26. X. J. Lu, S. J. Ding, T. Q. Lin, X. L. Mou, Z. L. Hong, and F. Q. Huang, “Ta2O5 nanowires: a novel synthetic method and their solar energy utilization,” Dalton Trans. 41(2), 622–627 (2012). [CrossRef]  

27. X. Ling, L. M. Xie, Y. Fang, H. Xu, H. L. Zhang, J. Kong, M. S. Dresselhaus, J. Zhang, and Z. F. Liu, “Can graphene be used as a substrate for Raman enhancement?” Nano Lett. 10(2), 553–561 (2010). [CrossRef]  

28. L. B. Gao, W. C. Ren, B. L. Liu, R. Saito, Z. S. Wu, S. S. Li, C. B. Jiang, F. Li, and H. M Cheng, “Surface and interference coenhanced Raman scattering of graphene,” ACS Nano 3(4), 933–939 (2009). [CrossRef]  

29. A. K. Geim and K. S. Novoselov, “The rise of graphene,” Nat. Mater. 6(3), 183–191 (2007). [CrossRef]  

30. R. R. Nair, P. Blake, A. N. Grigorenko, K. S. Novoselov, T. J. Booth, T. Stauber, N. M. R. Peres, and A. K. Geim, “Fine structure constant defines visual transparency of graphene,” Science 320(5881), 1308 (2008). [CrossRef]  

31. N. Jung, A. C. Crowther, N. Kim, P. Kim, and L. Brus, “Raman enhancement on graphene: adsorbed and intercalated molecular specie,” ACS Nano 4(11), 7005–7013 (2010). [CrossRef]  

32. W. Ji, B. Zhao, and Y. Ozakia, “Semiconductor materials in analytical applications of surface-enhanced Raman scattering,” J. Raman Spectrosc 47(1), 51–58 (2016). [CrossRef]  

33. I. Alessandri and J. R. Lombardi, “Enhanced Raman scattering with dielectrics. Chemical reviews,” Chem. Rev. 116(24), 14921–14981 (2016). [CrossRef]  

34. F. Zhang, H. R. Sun, X. Chen, J. Yang, L. Shi, T. Chen, Z. Y. Bao, J. Q. Liu, and Y. C Wu, “Highly efficient photoinduced enhanced Raman spectroscopy (PIERS) from plasmonic nanoparticles decorated 3D semiconductor arrays for ultrasensitive, portable, and eecyclable detection of organic pollutants,” ACS Sens. 4(6), 1670–1681 (2019). [CrossRef]  

35. S. Stankovich, D. A. Dikin, G. H. B. Dommett, K. M. Kohlhaas, E. J. Zimney, E. A. Stach, R. D. Piner, S. T. Nguyen, and R. S. Ruoff, “Graphene-based composite materials,” Nature 442(7100), 282–286 (2006). [CrossRef]  

36. K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, Y. Zhang, S. V. Dubonos, I. V. Grigorieva, and A. A. Firsov, “Electric field effect in atomically thin carbon films,” Science 306(5696), 666–669 (2004). [CrossRef]  

37. X. Li, J. Lin, and J. Zhao, “Preparation of a novel three-dimensional TiO2 macroporous/TiO2 nanorods/Ag nanoparticles composite nanostructure and its use as SERS substrates,” Nano 12(05), 1750052 (2017). [CrossRef]  

38. S. Q. Huang, C. J. Wu, Y. Y. Wang, X. Yang, R. Yuan, and Y. Q. Chai, “Ag/TiO2 nanocomposites as a novel SERS substrate for construction of sensitive biosensor,” Sensors and Actuators B: Chemical 339(15), 129843 (2021). [CrossRef]  

39. T. C. Gong, J. Zhang, Y. Zhu, X. Wang, X. Zhang, and J. Zhang, “Optical properties and surface-enhanced Raman scattering of hybrid structures with Ag nanoparticles and graphene,” Carbon 102, 245–254 (2016). [CrossRef]  

40. Y. Zhao, D. Yang, X. Y. Li, Y. Liu, X. Hu, D. F. Zhou, and Y. Lu, “Toward highly sensitive surface-enhanced Raman scattering: the design of a 3D hybrid system with monolayer graphene sandwiched between silver nanohole arrays and gold nanoparticles,” Nanoscale 9(3), 1087–1096 (2017). [CrossRef]  

41. F. Wei, Y. H. Lee, S. Pedireddy, Q. Zhang, T. X. Liu, and X. Y. Ling, “Graphene oxide and shape-controlled silver nanoparticle hybrids for ultrasensitive single-particle surface-enhanced Raman scattering (SERS) sensing,” Nanoscale 6(9), 4843–4851 (2014). [CrossRef]  

42. J. Huang, L. M. Zhang, B. Chen, N. Ji, F. H. Chen, Y. Zhang, and Z. J. Zhang, “Nanocomposites of size-controlled gold nanoparticles and graphene oxide: formation and applications in SERS and catalysis,” Nanoscale 2(12), 2733–2738 (2010). [CrossRef]  

43. Z. Li, S. Z. Jiang, Y. Y. Huo, T. Y. Ning, A. H. Liu, C. Zhang, Y. He, M. H. Wang, and C. H. Li, “3D silver nanoparticles with multilayer graphene oxide as a spacer for surface enhanced Raman spectroscopy analysis,” Nanoscale 10(13), 5897–5905 (2018). [CrossRef]  

44. C. Zhang, S. Z. Jiang, Y. Y. Huo, A. H. Liu, S. C. Xu, X. Y. Liu, Z. C. Sun, Y. Y. Xu, Z. Li, and B. Y. Man, “SERS detection of R6G based on a novel graphene oxide/silver nanoparticles/silicon pyramid arrays structure,” Opt. Express 23(19), 24811–24821 (2015). [CrossRef]  

45. L. Wang, Y. Zhang, Y. Q. Yang, and J. Zhang, “Strong dependence of surface enhanced Raman scattering on structure of graphene oxide film,” Materials 11(7), 1199 (2018). [CrossRef]  

46. X. L. Zhang, Y. S. Peng, Y. Yang, J. Chen, H. W. Lian, X. Q. Cui, and W. T. Zheng, “Hydrothermal reduction of graphene oxide; effect on surface-enhanced Raman scattering,” J Raman Spectrosc 48(1), 97–103 (2017). [CrossRef]  

47. H. P. Yang, H. L. Hu, Z. H. Ni, C. K. Poh, C. X. Cong, J. Y. Lin, and T. Yu, “Comparison of surface-enhanced Raman scattering on graphene oxide, reduced graphene oxide and graphene surfaces,” Carbon 62, 422–429 (2013). [CrossRef]  

48. Q. Q. Ding, J. Wang, X. Y. Chen, H. Liu, Q. J. Li, Y. L. Wang, and S. K. Yang, “Quantitative and sensitive SERS platform with analyte enrichment and filtration function,” Nano Lett. 20(10), 7304–7312 (2020). [CrossRef]  

49. H. Y. Chen, M. H. Lin, C. Y. Wang, Y. M. Chang, and S. J. Gwo, “Large-scale hot spot engineering for quantitative SERS at the single-molecule scale,” J. Am. Chem. Soc. 137(42), 13698–13705 (2015). [CrossRef]  

50. W. S. Hummers and R. E. Offeman, “Preparation of graphitic oxide,” J. Am. Chem. Soc. 80(6), 1339 (1958). [CrossRef]  

51. Y. G. Sun and Y. N. Xia, “Mechanistic study on the replacement reaction between silver nanostructures and chloroauric acid in aqueous medium,” J. Am. Chem. Soc. 126(12), 3892–3901 (2004). [CrossRef]  

52. R. X. He, R. Liang, P. Peng, and Y. N. Zhou, “Effect of the size of silver nanoparticles on SERS signal enhancement,” J. Nanopart Res 19(8), 267 (2017). [CrossRef]  

53. D. C. Elias, R. R. Nair, T. M. G. Mohiuddin, S. V. Morozov, P. Blake, M. P. Halsall, A. C. Ferrari, D. W. Boukhvalov, M. I. Katsnelson, A. K. Geim, and K. S. Novoselov, “Control of graphene’s properties by reversible hydrogenation: evidence for graphane,” Science 323(5914), 610–613 (2009). [CrossRef]  

54. S. J. Lei, C. J. Tao, J. L. Li, X. Zhao, and W. Z. Wang, “Visible light-induced charge transfer to improve sensitive surface-enhanced Raman scattering of ZnO/Ag nanorod arrays,” Applied Surface Science 452, 148–154 (2018). [CrossRef]  

55. S. Chakraborty, S. Chowdhury, and P. D. Saha, “Adsorption of Crystal Violet from aqueous solution onto NaOH-modified rice husk,” Carbohydr. Polym. 86(4), 1533–1541 (2011). [CrossRef]  

56. E. C. L. Ru, E. Blackie, M. Meyer, and P. G. Etchegoin, “Surface enhanced Raman scattering enhancement factors: a comprehensive study,” J. Phys. Chem. C 111(37), 13794–13803 (2007). [CrossRef]  

57. A. R. Shalaby, W. H. Emam, and M.M. Anwar, “Mini-column assay for rapid detection of malachite green in fish,” Food Chemistry 226, 8–13 (2017). [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 (7)

Fig. 1.
Fig. 1. Preparation process of GO/Ag/Ta2O5 composite SERS substrate.
Fig. 2.
Fig. 2. (a-d) SEM image of Ta2O5, AgNPs, Ag/Ta2O5 and GO/Ag/Ta2O5 composite substrates. (e) SEM image of Ag/Ta2O5 and GO/Ag/Ta2O5 composite substrates. (f) The energy dispersion spectrometer of the GO/Ag/Ta2O5 composite substrates. (g) EDS elemental maps from the GO/Ag/Ta2O5 composite substrates. (h,i) The size distribution of Ta2O5 and AgNPs. (j) The UV-Vis absorption spectrum of AgNPs. (k) The Raman spectrum of GO.
Fig. 3.
Fig. 3. (a,b) Ta4f and O1s detailed XPS spectra of purchased Ta2O5 (c) Ag3d detailed XPS spectra of Ag/Ta2O5 composite substrate. (d,e) Ta4f and O1s detailed XPS spectra of Ag/Ta2O5 composite substrate.(f) Survey XPS spectra of Ag/Ta2O5 composite substrate.
Fig. 4.
Fig. 4. (a) The Raman spectra of different concentrations of R6G on SiO2, Ta2O5, AgNPs and Ag/Ta2O5 composite substrates. (b) Raman spectra of R6G on the Ag/Ta2O5 composite substrates from 10−7 M to 10−12 M. (c) Log−log plot of average intensity of SERS signals at 613 and 1363 cm−1 versus the concentration of R6G. (d) The intensity distributions of the 613, 1364 and 1511 cm−1 peaks for R6G. (e) Corresponding Raman mapping of the 613 cm−1 characteristic peak of R6G (10−7 M). (f) The charge transfer between R6G, Ta2O5 and AgNPs.
Fig. 5.
Fig. 5. (a) The Raman spectra of different concentrations of CV on SiO2, Ag/Ta2O5 and GO/Ag/Ta2O5 composite substrates. (b) Raman spectra of CV on the GO/Ag/Ta2O5 composite substrates from 10−6 M to 10−11 M. (c) Log−log plot of average intensity of SERS signals at 915 and 1621 cm−1 versus the concentration of CV. (d) Raman spectra of CV (10−7 M) on GO/Ag/Ta2O5 composite substrates at different time. (e) Raman spectra of CV (915 cm-1) on GO/Ag/Ta2O5 composite substrate and Ag/Ta2O5 composite substrate. (f) Corresponding Raman mapping of the 915 cm−1 characteristic peak of CV (10−6 M).
Fig. 6.
Fig. 6. (a) The Raman spectra of the measured lake water without MG on SiO2, the Raman spectra of 10−2 M MG prepared with lake water on SiO2, and the Raman spectra of 10−10 M MG droplets prepared with lake water on the composite substrate. (b) Raman spectra of MG on the GO/Ag/Ta2O5 composite substrates from 10−5 M to 10−10 M. (c) Log−log plot of average intensity of SERS signals at 1618 and 1368 cm−1 versus the concentration of MG.
Fig. 7.
Fig. 7. (a)-(d) Simulated electric field distribution of Ta2O5, AgNPs, Ag/Ta2O5 and GO/Ag/Ta2O5 composite substrates.

Tables (1)

Tables Icon

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

Equations (2)

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

R S D = S x ¯ × 100 % = i = 1 n ( x i x ¯ ) 2 n 1 x ¯ × 100 % .
E F = I S E R S / N S E R S I R S / N R S
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.