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

Noble metal modified ReS2 nanocavity for surface-enhanced Raman spectroscopy (SERS) analysis

Open Access Open Access

Abstract

The rhenium disulphide (ReS2) nanocavity-based surface enhanced Raman scattering (SERS) substrates ware fabricated on the gold-modified silicon pyramid (PSi) by thermal evaporation technology and hydrothermal method. In this work, the ReS2 nanocavity was firstly combined with metal nanostructures in order to improve the SERS properties of ReS2 materials, and the SERS response of the composite structure exhibits excellent performance in sensitivity, uniformity and repeatability. Numerical simulation reveals the synergistic effect of the ReS2 nanocavity and the plasmon resonance generated by the metal nanostructures. And the charge transfer between the metal, ReS2 and the analytes was also verified and plays an non-ignorable role. Besides, the plasmon-driven reaction for p-nitrothiophenol (PNTP) to p,p'-dimercaptobenzene (DMAB) conversion was successfully in-situ monitored. Most importantly, it is found for the first time that the SERS properties of ReS2 nanocavity-based substrates are strongly temperature dependent, and the SERS effect achieves the best performance at 45 °C. In addition, the low concentration detection of malachite green (MG) and crystal violet (CV) molecules in lake water shows its development potential in practical application.

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

1. Introduction

As an interference-free ultrasensitive analytical technique, surface-enhanced Raman spectroscopy (SERS) has been widely used in various biochemical assays at the single-molecule level [13]. As we all know, generally, the enhancement mechanism of SERS signal of adsorbed probe molecule includes electromagnetic mechanism (EM) and chemical mechanism (CM) [4,5]. Furthermore, EM can be ascribed to enhanced local electromagnetic fields (usually called hot spots) at the surface of coinage metal nanostructures, and CM is related to charge transfer or energy transfer between the analytes and the substrate. In addition, Electromagnetic enhancements caused by the light-excited surface plasmon resonance (SPR) are usually orders of magnitude higher than chemical enhancements caused by charge transfer or energy transfer process [610]. However, the development of SERS substrates composed solely of noble metals were limited due to the disadvantages of high cost and poor biocompatibility [11].

Simultaneously, semiconductor SERS substrates have also been widely developed [12,13]. Transition metal dichalcogenides (TMDs), which have already attracted the intensive interest of many researchers, have broad application prospects because of their distinctive optical and electrical properties [1419]. Especially, TMDs play an important role in exploring the CM of SERS. As a new member of the TMDs family, Rhenium disulphide (ReS2) possesses many extraordinary properties different from traditional TMDs materials such as MoS2, MoSe2, WS2 and WSe2, because of its distorted octahedral (1T) crystal structure [2023]. ReS2 maintains a direct band gap semiconductor owing to electronic and vibrational decoupling, whether it is single-layer, few-layer or bulk. As a consequence, ReS2 possess huge development potential and is high-profile in the field of catalysis, energy storage, energy harvesting and solid-state electronics in recent years. Besides, Peng Miao et al. also reported that ReS2 possesses the SERS effect due to the charge transfer process between ReS2 and the analytes [24]. Nevertheless, the Raman enhancement effect of ReS2 is greatly inferior to that of the noble metal nanostructures, which limits its practical applications for the detection of aromatic molecules, toxic substances and biomolecules, etc. Therefore, it is of great significance to optimize the SERS effect of ReS2. Jihyung Seo et al. optimized the electronic structure of ReS2 and adjusted the surface polarity by doping different concentrations of oxygen to improve the dipole-dipole interactions and charge-transfer resonance between the ReS2 and the analytes. Moreover, the femtomolar detection level of R6G was achieved by constructing a graphene/ReOxSy vertical heterostructure [25].

Though, these works promoted the application of ReS2 in SERS, integrating the ReS2 with metal nanostructures has not been investigated yet. Furthermore, compared with the SERS substrate with two-dimensional planar hot spots, the SERS substrates with three-dimensional and cavity structure possess its unique advantages. Zewen Zuo et al. developed a kind of silver conical cavity array [26]. The highly sensitive SERS substrate was obtained by effectively coupling the surface plasmons of nanoparticles with the cavity mode. The cavity structure can not only increase the surface area contributing to molecular adsorption, but also expand the SERS region from the surface area to the hollow area within the nanocavity [27]. The concept of volume-enhanced Raman scattering (VERS) was proposed by Xingang Zhang et al. And their hollow nanocones substrate greatly improved the Raman signal of single virus [28].

Currently, combining two materials with different properties into hybrid structure as SERS substrate is an effective method to improve the overall Raman enhancement performance [2933]. In this work, ReS2 nanocavity (ReS2NC) decorated with silver nanoparticles (AgNP) is firstly synthesized on pyramid silicon (PSi) modified by gold nanoparticles (AuNP) using hydrothermal technique and thermal evaporation technology. By simply adjusting the thickness of silver, the SERS performance of the substrate (called PSi/Au/ReS2NC/Ag) is optimized. Benefiting from large surface area and dense hot spots, PSi/Au/ReS2NC/Ag shows excellent detection performance for dye molecules and can be used to in-situ monitor the conversion of p-nitrothiophenol (PNTP) to p,p'-dimercaptobenzene (DMAB) driven by plasmon. In addition, the temperature-responsive behavior of the substrates in SERS were first investigated. We believe this research will promote the application of metal-modified semiconductor substrates in SERS and deepen our understanding of its underlying properties.

2. Experimental

2.1 Fabrication of the PSi/Au/ReS2NC/Ag

The preparation procedures of the PSi/Au/ReS2NC/Ag are schematically shown in Fig. 1. Simply put, Au and Ag are deposited on the substrate by vacuum thermal evaporation, and ReS2 is grown on the substrate by hydrothermal method. As we have reported previously, pyramid silicon (PSi) was fabricated by wet etching technology. In a typical experiment, several pieces of PSi samples were ultrasonically washed by acetone, ethanol and deionized water (DI water) for 15 min separately. The gold film with 7 nm thickness was deposited on clean PSi using thermal evaporation method, the sample was then put in a quartz tube and annealed for 30 min at 700°C. 26.8 mg of ammonium perrhenate (NH4ReO4), 20.8 mg of hydroxylamine hydrochloride (NH2OH·HCl) and 36.3 mg of L-cysteine ware ultrasonically dispersed in 10 ml ultrapure water (UP water) well, and then transferred to a 50 ml polytetrafluoroethylene-lined autoclave, followed by the PSi modified by AuNPs was put into the above autoclave, sealed and stored at 200°C with different reaction time of 2 h, 4 h and 6 h. Let the sample cool to room temperature naturally, take it out and rinse it well with UP water. Afterwards, the silver was deposited on prepared PSi/Au/ReS2NC by vacuum thermal evaporation technology. In addition, by adjusting the thermal deposition thickness from 5 nm to 25 nm, the PSi/Au/ReS2NC/Ag with the optimal SERS performance was obtained.

 figure: Fig. 1.

Fig. 1. Schematic diagram of the fabrication of PSi/Au/ReS2NC/Ag.

Download Full Size | PDF

2.2 Characterization

Scanning electron microscope (SEM, Zeiss Gemini Ultra-55), transmission electron microscope (TEM JEM-2100) and energy dispersive spectrometer (EDS) were used to characterize the morphology of the fabricated samples. The content of elements was also measured by X-ray photoelectron spectroscopy (XPS Thermo Scientific Escalab 250Xi). The SERS spectrum was collected by the Raman spectrometer (Horiba HR Evolution) with the excited laser of 532 nm to study the SERS performance of the prepared PSi/Au/ReS2NC and PSi/Au/ReS2NC/Ag SERS substrates. The 50 × objective lens was used. The integration time and the diffraction grating were set to 4 s and 600 gr/nm respectively. For the plasmon-driven, surface-catalyzed reactions, the PSi/Au/ReS2NC and PSi/Au/ReS2NC/Ag SERS substrates were immersed in 10−3 M PNTP solution for more than 1 h. Then, the samples were rinsed with ethanol several times and dried in N2 condition before the Raman test. A temperature controller equipped with a liquid nitrogen tank was used to controllably adjust the temperature of the SERS substrates. The temperature change rate was set to 50 °C per minute. SERS data was collected every 10 °C change.

3. Results and discussion

3.1 Characterization of the PSi/Au/ReS2NC/Ag structure

A mass of tightly arranged AuNP ware uniformly attached to the PSi surface (Supplement 1, Fig. S1A). The histograms of the size distribution of AuNP diameter with average 30 nm and interparticle gap with average 11 nm are also shown in Supplement 1, Figs. S1B and S1C respectively. From Fig. 2(A), we can see clearly that ReS2 layer has been grown successfully along the PSi/Au surface. ReS2 nanosheets with 10 nm wall thickness were vertically overlaid on the AuNP to form dense nanocavities with 60 nm cavity width. The inset in Fig. 2(A) show boundary of the PSi/Au coated or uncoated ReS2 nanosheets. Our experiments indicated that the existence of AuNP has a positive effect on the growth of ReS2 on PSi [34]. As shown in Supplement 1, Fig. S2, ReS2 nanocavity is difficult to directly grow on PSi surface without AuNP, and form nanospheres made up of nanosheets scattered sporadically in the valley of PSi, which is attributed to the weak interaction between ReS2 and Si. Figure 2(B) shows the SEM image of the PSi/Au/ReS2NC/Ag, and the illustration in Fig. 2(B) exhibits the side view of Au/ReS2NC/Ag where its thickness of is about 70 nm. To further confirm the successful preparation of PSi/Au/ReS2NC/Ag hybrid, HRTEM was also carried as shown in Fig. 2(C). It could be clearly seen that AuNP marked with yellow line is attached to ReS2 nanosheets, and the inter-layer spacing of 0.61 and 0.24 nm correspond to the (100) plane of ReS2 nanosheets and the (111) plane of AuNP, separately. In addition, the Ag with 0.23 nm inter-layer spacing in agreement with the (111) plane is distributed at the interface of ReS2 nanosheets as presented in Fig. 2(D). It is worth noting that the AuNP is located in the center of ReS2 nanosheets and the Ag located in the edge of ReS2 nanosheets, which is consistent well with our experimental expectations. Moreover, as shown in Fig. 2(E), the elements composition of the PSi/Au/ReS2NC/Ag was also measured by EDS elemental mappings, which clearly shows the presence of Au (cyan), Re (orange), S (green) and Ag (purplish red). In order to make a comparison, Si element mapping was also put into Fig. 2(E). Si element is mainly distributed in the valley of pyramid, while Au, Re, S and Ag elements are mainly distributed in the inclined plane and top of pyramid, which is related to the sparse distribution density of AuNP in the valley areas. This phenomenon is consistent with the experimental results discussed previously.

 figure: Fig. 2.

Fig. 2. (A) SEM image of PSi/Au/ReS2NC. (B) SEM image of PSi/Au/ReS2NC/Ag. (C) and (D) HRTEM image of PSi/Au/ReS2NC/Ag. (E) EDS elemental mappings from Au, Re, S, Ag and Si on the PSi/Au/ReS2NC/Ag.

Download Full Size | PDF

The elemental constituent and chemical states of the fabricated substrates ware confirmed by using X-ray photoelectron spectroscopy (XPS). It can be clearly observed that the two characteristic peaks of PSi/Au/ReS2NC substrate are 42.47 and 44.85 eV in Fig. 3(A), corresponding to the level peaks of Re4+ 4f7/2 and Re4+ 4f5/2, respectively. It is interesting to note that the binding energies of 4f7/2 and 4f5/2 peaks of Re4+ in PSi/Au/ReS2NC/Ag are negative shifted by 0.1 eV. In addition, the two Ag 3d peaks at 368.79 and 374.80 eV in the PSi/Au/ReS2NC/Ag are also observed as shown in Fig. 3(B), corresponding to the level peaks of 3d5/2 and 3d3/2, separately. However, compared with the bulk Ag, the binding energies of the two Ag 3d have positively shifts around 0.29 eV [35]. The above results indicate that the electrons transfer from Ag to ReS2 occurs. The existence of charge transfer between ReS2 and Ag is suitable for improving the chemical enhancement and catalytic capability of SERS substrate. The high-resolution S 2p spectrum of the PSi/Au/ReS2NC and PSi/Au/ReS2NC/Ag substrate are shown in Supplement 1, Fig. S3, and the atomic percentage of Re and S is close to 1:2 (Supplement 1, Fig. S4), which confirms the elemental constituent and chemical states of the ReS2. Figure 3(C) depicts the Raman signals of ReS2 from the PSi/Au/ReS2NC and PSi/Au/ReS2NC/Ag. Because of its low symmetry, the Raman spectrum of ReS2 display more characteristic peaks than other TMDs. The basic Raman modes (E2g, A1g and E1g) are coupled with each other and acoustic phonons. The Raman bands at 157, 163 and 316 cm-1 (Red curves) refer to the in-plane vibration, while the Raman band at 213 cm-1 (Green curve) refers to the mostly out-of-plane vibration. The scanning Raman mapping of E2g mode of the PSi/Au/ReS2NC is also exhibited in Fig. 3(D). The low color difference indicates well uniformity of ReS2.

 figure: Fig. 3.

Fig. 3. (A) XPS spectra of Re of PSi/Au/ReS2NC and PSi/Au/ReS2NC/Ag sample. (B) XPS spectra of Ag of PSi/Au/ReS2NC/Ag sample. (C) The Raman spectra of PSi/Au/ReS2NC and PSi/Au/ReS2NC/Ag sample. (D) The scanning Raman mapping of E2g mode of the PSi/Au/ReS2NC sample.

Download Full Size | PDF

3.2 Optimization of SERS performance of the PSi/Au/ReS2NC/Ag

As shown in Supplement 1, Fig. S5, Ag of changing from 5 to 25 nm was steamed onto the PSi/Au/ReS2NC to optimize the SERS response of the PSi/Au/ReS2NC/Ag. An enlarged view of the PSi/Au/ReS2NC is displayed in Supplement 1, Fig. S5A. When the deposition thickness is 5 nm, the Ag distribut granularly and sparsely on the ReS2 nanosheets (Supplement 1, Fig. S5B). When the deposition thickness reaches 10 nm, the diameter of the AgNP becomes larger (Supplement 1, Fig. S5C). By further increasing the deposition thickness to 15 nm, the diameter of the AgNP continues to increase, and they are densely and uniformly adsorbed on the ReS2 nanosheets (Supplement 1, Fig. S5D). As the deposition thickness increased to 20 nm, AgNP began to contact each other on the ReS2 nanosheets to form a continuous film, and many protrusions were produced at the same time (Supplement 1, Fig. S5E). The Ag film becomes thicker and completely and uniformly covers the ReS2 nanosheet when the deposition thickness continues to increase to 25 nm (Supplement 1, Fig. S5F).

In order to assess the SERS performance of the above substrates, Rhodamine 6G (R6G) was chosen as the analytes to conduct SERS contrast detection. The observed vibrational bands from Fig. 4(A) are well matched with the previously reported works. The peak at 614 cm-1 is attributed to the in-plane vibration of C-C-C deformation, the peak at 776 cm-1 refers to the out-of-plane vibration of C-H bond deformation, and the peak at 1650 cm-1 applies to the aromatic C-C stretching [36]. Figure 4(B) demonstrates the line graph of the peak intensities of the characteristic bands at 614 and 776 cm-1 with the Ag thickness from 5 nm to 25 nm. With the increase of Ag thickness, the SERS signals of R6G molecules (10−5 M) was enhanced gradually and reached maximum at the 15 nm thickness. Note that when the Ag thickness continued to increase, the SERS signal strength declined instead. This is because the over-thick Ag completely covers the ReS2 layer to form a flat film, resulting in a reduction in the density of hot spots.

 figure: Fig. 4.

Fig. 4. (A) SERS spectrum of R6G (10−5 M) collected on the PSi/Au/ReS2NC/Ag with varying deposition thickness of Ag. (B) The broken line graphs of SERS intensities at 614 and 776 cm-1 with Ag deposition thickness. (C)-(G) Simulated vertical electric field distributions of the ReS2NC deposited different Ag thickness of (C) 5 nm, (D) 10 nm, (E) 15 nm, (F) 20 nm, (G) 25 nm. (H) The simulated enhancement factor (|E|2/|E0|2) of the different hybrid structures.

Download Full Size | PDF

3.3 SERS enhanced mechanism of the PSi/Au/ReS2NC/Ag system

To deeply understand the variation of the SERS intensity changes with the Ag thickness, the electric field distribution of different SERS substrates was simulated by commercial software COMSOL. Due to the thick ReS2 layer, the electric field enhancement generated by the AuNP was ignored in the simulation. The refractive index of ReS2 was set as n(ReS2) = 3.99 + 0.17i under 532 nm laser irradiation [37]. According to the SEM images, the thickness of the ReS2 nanosheets was set as 10 nm, the interval between the nanosheets was set as 60 nm, the diameter of the AgNP was set as 20 nm, 22 nm, 25 nm, and the thickness of the Ag film was set as 14 and 16 nm, respectively. As the particle size increases and the gap between particles decreases, the electric field intensity increases significantly, and reaches a maximum at 15 nm (Fig. 4(C)-(E)). The electric field is primarily distributed in the cavity and around the nanoparticles because of the coupling of the surface plasmons of the AgNP and the cavity mode. However, the number of hot spots is drastically reduced due to the formation of the silver film, resulting in a decrease in electric field intensity (Fig. 4(F) and (G)). Figure 4(H) shows the change trend of the enhancement factors (|E|2/|E0|2) of the above five samples, which is consistent with the previous experimental results. Besides, ReS2 nanosheets can be used as a superior adsorbent for analytes, which is helpful to improve the SERS response of the PSi/Au/ReS2NC/Ag.

Apart from electromagnetic enhancement, the chemical enhancement of ReS2 caused by charge transfer should also be considered in our samples as shown Fig. 5. It was reported that the conduction band minimum (CBM) and valence band maximum (VBM) of monolayer ReS2 are -6.15 and -4.26 eV, respectively [17]. And the highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO) levels of R6G are -5.7 and -3.4 eV, separately [38]. The photoexcited electrons induced by surface plasmon resonance of AgNP were injected into the CB of ReS2 when the visible laser was irradiated on the substrate, then transferred to LUMO of the probe molecules. Furthermore, from an energy point of view, the establishment of Schottky barrier between AgNP and ReS2 facilitates the transport of the photoexcited electrons from AgNP to ReS2.

 figure: Fig. 5.

Fig. 5. Schematic diagram of the proposed charge transfer mechanism.

Download Full Size | PDF

3.4 Surface catalytic reaction performance of the composite structure

Some chemical reactions can be driven by hot electrons induced by SPR on the surface of coinage metal nanostructures [3943]. Up to present, the catalytic reactions driven by plasmon have been extremely critical in environmental protection, solar energy, and chemical industry for low-energy requirements and high efficiency [44,45]. Under normal circumstances, the extremely short hot electron lifetime greatly bounds the efficiency of hot electron-induced reduction reactions owing to the fast decay of SPR. The combination of metals and semiconductors materials is an effective method to transfer and extend the lifetime of hot electrons. Therefore, the PSi/Au/ReS2NC/Ag SERS substrates were used to in-situ monitor the reduction reaction from PNTP to DMAB and verify the catalytic ability. The SERS spectra of Ag free substrate only shows the 1340 cm-1 peak, representing the ${V_{({N{O_2}} )\; }}$ of PNTP, which is related to the thicker ReS2 layer, as shown in Fig. 6(A). For the sample with a deposition thickness of 5 nm of silver, the peaks at 1397 and 1445 cm-1 representing the ${V_{({N = N} )\; }}$ of DMAB molecules appeared after 10 seconds of reaction (Fig. 6(B)). However, the intensity of the DMAB band is weak compared to the intensity of the PNTP band, indicating poor conversion efficiency. Compared with the ${V_{({N{O_2}} )\; }}$ band, the intensity of the ${V_{({N = N} )\; }}$ bands are much stronger after the reaction for 10 s when the deposited thickness of Ag increased to 10 nm or more, indicating high conversion efficiency (Fig. 6(C)-(F)). As shown in Fig. 6(G), the catalytic efficiency of the substrates was evaluated using the time-dependent formula of ${I_{{V_{({N = N} )}}}}/{I_{{V_{({N{O_2}} )}}}}$, where ${I_{{V_{({N = N} )}}}}$ represents the SERS intensity of DMAB band at 1445 cm-1, while ${I_{{V_{({N{O_2}} )}}}}$ represents the SERS intensity of PNTP band at 1340 cm-1. The results show that the reaction rate is the fastest when the deposited thickness of Ag reaches 15 nm. It is worth noting that the PSi/Au/ReS2NC/Ag with higher intensity of hot spots is expected to generate more “hot electrons” to accelerate this reaction. However, too much Ag deposited on the ReS2 nanosheets would act as a barrier layer which prevents the molecules from contacting ReS2, resulting the electron and hole recombination more easily, and ultimately reducing the reaction efficiency. In addition, the SERS strength of ${V_{({N = N} )\; }}$ band at 1445 cm-1 after 60 s reaction collected from PSi/Au/ReS2NC/Ag with different deposited thickness of Ag are exhibited in Fig. 6(H). This change trend is consistent with the SERS spectrum of R6G and the simulated results.

 figure: Fig. 6.

Fig. 6. (A)-(F) In-situ SERS monitoring the dimerization of PNTP to DMAB for PSi/Au/ReS2NC/Ag with different deposited thickness of Ag. (G) Time dependence of ${I_{{V_{({N = N} )}}}}/{I_{{V_{({N{O_2}} )}}}}$ for PSi/Au/ReS2NC/Ag with different Ag thickness under 532 nm laser excitation. (H) Ag thickness-dependent SERS intensities of the ${V_{({N = N} )\; }}$ band.

Download Full Size | PDF

3.5 Investigation of SERS activity, signal homogeneity and substrate repeatability

For the sake of detection of the SERS sensitivity of the PSi/Au/ReS2NC/Ag with 15 nm Ag thickness, the SERS spectrum of R6G aqueous solution with changing concentrations (10−5 - 10−11 M) was measured (Fig. 7(A)). Although the SERS intensity of R6G was reduced with the decreasing concentration, SERS signal of R6G could still be detected when the R6G concentration is as low as 10−11 M. Besides, as shown in Fig. 7(B), Raman characteristic peaks of 614 and 776 cm-1 were chosen for linear fitting to study the correlation between SERS signal intensity and R6G concentration. The value of fitting correlation coefficient (R2) of 614 and 776 cm-1 reached 0.986. The fitting formula of intensity change with concentration is as follows: $\textrm{Log}\; I = A\; \textrm{Log}\; C + B$, where I is the SERS intensity of the probe molecules, A refers to the slope, C refers to the concentrations and B refers to the intercept. For 614 and 776 cm-1 peaks, the fitting formulas are $\textrm{Log}\; I = 0.358\; \textrm{Log}\; C + 5.901$ and $\textrm{Log}\; I = 0.360\; \textrm{Log}\; C + 5.510$, respectively. Excellent linear relationship under the log scale enables quantitative detection of dye molecules. As a comparison, the SERS spectrum of R6G aqueous solution with various concentrations (10−4 - 10−7 M) were also measured in PSi/Au/ReS2NC (Supplement 1, Fig. S6). Even if gold does not play a role in the direct Raman enhancement, the substrate still shows a good ability to detect low concentration molecules. In addition to the charge transfer effect, the nanocavity has a unique cavity structure, which can effectively capture light by concentrated electric field and improve the light capture ability of the substrate. The decrease of detection line in PSi/Au/ReS2NC/Ag could be explained by the synergistic effect of physical enhancement and chemical enhancement, as mentioned above. Figure 7(C) shows the Raman mapping of R6G molecules at 614 cm-1 from the same sample. The relatively uniform color distribution in maps indicates that the homogeneity of the sample is relatively good. In addition to sensitivity and uniformity, repeatability is also an indispensable indicator of the performance of SERS substrates. Figure 7(D) shows the SERS spectrum of R6G collected from ten samples obtained using the same preparation method. For the sake of clarity, as shown in Fig. 7(E) and (F), the 614 and 776 cm-1 peaks were chosen to study the repeatability of SERS spectra, and the relative standard deviation (RSD) could be calculated using the followed equation:

$$RSD = \frac{{\sqrt {\frac{{\mathop \sum \nolimits_{i = 1}^n {{({{I_i} - \bar{I}} )}^2}}}{{n - 1}}} }}{{\bar{I}}},$$
where ${I_i}$ refers to the intensity of R6G SERS signal collected on different samples, $\bar{I}$ refers to the average intensity of the same band in entire the SERS spectrum collected on different samples, and n refers to the number of samples ($n$=10). The$\; RSD\; $was calculated to be 7.27% at 614 cm-1 and 8.07% at 776 cm-1, which indicates the superior repeatability of the PSi/Au/ReS2NC/Ag.

 figure: Fig. 7.

Fig. 7. (A) SERS spectrum of R6G (10−5 - 10−11 M) in PSi/Au/ReS2NC/Ag. (B) The relation between the intensity of SERS signals and the concentration of R6G at 614 and 776 cm−1 in logarithmic coordinates. (C) The scanning Raman mappings of characteristic peak of R6G (10−6 M) at 614 cm-1. (D) SERS spectrum of R6G (10−6 M) collected on different substrates. SERS intensity at 614 cm-1 (E) and 776 cm-1 (F) obtained from Raman spectra in (D).

Download Full Size | PDF

3.6 Temperature response behavior of SERS properties of composite substrates

The regulation of hot spots is an important research direction in the field of SERS. Previous studies have shown that the morphology of ReS2 nanosheets changes with ambient temperature [46]. For the extremely weak interlayer interaction of ReS2, out-of-plane bending would occur in the ReS2 layers and the bending could recover to initial state as the temperature drops. In this work, the SERS signals of PSi/Au/ReS2NC and PSi/Au/ReS2NC/Ag were investigated by changing the test temperature. As shown in Fig. 8(A) and (B), the SERS signals collected from the PSi/Au/ReS2NC gradually decreased when the temperature was elevated from room temperature to 85 °C. While for the PSi/Au/ReS2NC/Ag, the SERS intensity gradually increased when the temperature rose from room temperature and reached maximum at 45 °C shown in Fig. 8(C) and (D). Subsequently, the SERS intensity decreased as the temperature increase to 85 °C. Interestingly, the SERS intensity increased again when the temperature decreased, and also reached the maximum when the temperature decreased to 45 °C. This experimental phenomenon can be expounded by the out-of-plane bending of ReS2 nanosheets with temperature changes as shown in Fig. 8(E). When the temperature rises, the ReS2 nanosheets begin to tilt from the vertical state. For the PSi/Au/ReS2NC, the analyte molecules absorbed on the bottom of the cavity could not be detected resulting in continuously decreased SERS signals as the temperature increases. While for the PSi/Au/ReS2NC/Ag, the gap among nanosheets would decrease as the temperature increases, which is beneficial for the elevation of the hot spot intensity bringing advantages to the SERS. When the temperature further heightens, the nanocavities would disappear, and thus resulting in a weak SERS response.

 figure: Fig. 8.

Fig. 8. (A) SERS spectrum of R6G (10−5) in PSi/Au/ReS2NC at 25 - 85 °C. (B) The broken line graphs of SERS signal intensity at 614 and 776 cm-1 with temperature in PSi/Au/ReS2NC. (C) SERS spectrum of R6G (10−5) in PSi/Au/ReS2NC/Ag at 25 - 85 °C. (D) The broken line graphs of SERS signal intensity at 614 and 776 cm-1 with temperature in PSi/Au/ReS2NC/Ag. (E) Schematic diagram of the morphology of ReS2 nanosheets changing with the temperature.

Download Full Size | PDF

3.7 Verification of detection ability for real samples of the PSi/Au/ReS2NC/Ag

Malachite Green (MG), a toxic chemical of triphenylmethane, is an effective fungicide, which has been widely used in aquaculture. However, chronic overuse of MG is harmful to human health and has carcinogenic effect. Therefore, many countries have banned the use of MG in the field of pollution-free aquaculture. In order to verify practicability of PSi/Au/ReS2NC/Ag substrate, we collected water from the local lake and used it as a solvent to prepare MG solutions with different concentrations. As shown in Fig. 9(A), the detection limit is as low as 10−9 M, reaching European and American standards. Similarly, the Raman characteristic peak at 1620 cm-1 is chosen to further survey the relationship between signal intensity and concentration (Fig. 9(B)). The fitting formula is $\textrm{Log}\; I = 0.385\; \textrm{Log}\; C + 6.131$, and R2 reaches 0.985, which could be used for quantitative detection of real samples. As a comparison, SERS signals of MG (10−8 M) in two different solvent of DI water and lake water were collected as shown in inset of Fig. 9(B). The SERS intensity of MG in lake water is weaker than that in DI water, which may be due to a large number of other impurities weakening the signal strength. Crystal Violet (CV), another effective bactericide, also has carcinogenic effect and is forbidden to be used in edible products. In likewise, lake water was also used as solvent to prepare CV solution for SERS measurement. Figure 9(C) exhibits the SERS spectrum of CV at different concentration. The fitting formula of intensity changing with concentration at 915 cm-1 is $\textrm{Log}\; I = 0.461\; \textrm{Log}\; C + 6.352$. And the R2 is 0.992, which can realize quantitative detection (Fig. 9(D)). As shown in the inset of Fig. 9(D), we again compared the SERS spectra of CV with the same concentration (10−8 M) in lake water and DI water. The result is similar to that of MG. Therefore, PSi/Au/ReS2NC/Ag sensor shows a potential ability to detect harmful substances in aquatic products.

 figure: Fig. 9.

Fig. 9. (A) SERS spectrum of MG (10−4 - 10−9 M) in lake water solution collected on PSi/Au/ReS2NC/Ag. (B) The relation between the intensity of SERS signals and the concentration of MG at 1620 cm−1 in logarithmic coordinates. (C) SERS spectrum of CV (10−6 - 10−10 M) in lake water solution collected on PSi/Au/ReS2NC/Ag. (D) The relation between the intensity of SERS signals and the concentration of CV at 915 cm−1 in logarithmic coordinates.

Download Full Size | PDF

4. Conclusions

In short, we developed a three-dimensional SERS substrate based on ReS2 nanocavity by using thermal evaporation and hydrothermal synthesis technology, and studied the SERS enhancement mechanism in detail, including physical and chemical mechanisms. The PSi/Au/ReS2NC/Ag exhibits good sensitivity, uniformity and repeatability for detection of R6G, and can be used to in-situ monitor the reaction for PNTP to DMAB conversion driven by plasmon. Moreover, in this study, it is found for the first time that the SERS performance of ReS2 nanocavity-based substrates show a strong temperature correlation, and the SERS effect achieve the best performance at 45 °C. We believe this result can be used as a new thought for the designing 2D materials/metal composite structures used in SERS effect with putting the thermal deformation as a vital factor. For actual application, MG and CV molecules in lake water were successfully detected, showing its development potential in the field of food safety.

Funding

National Natural Science Foundation of China (11774208, 11804200, 11904214, 11974222, 12004226); Natural Science Foundation of Shandong Province (ZR2020QA075); China Postdoctoral Science Foundation (2019M662423).

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.

Supplemental document

See Supplement 1 for supporting content.

References

1. 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,” Sens. Actuators, B 258, 163–171 (2018). [CrossRef]  

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

3. H. Y. Zhao, J. Jin, W. J. Tian, R. Li, Z. Yu, W. Song, Q. Cong, B. Zhao, and Y. Ozaki, “Three-dimensional superhydrophobic surface-enhanced Raman spectroscopy substrate for sensitive detection of pollutants in real environments,” J. Mater. Chem. A 3(8), 4330–4337 (2015). [CrossRef]  

4. C. Zhan, X. J. Chen, J. Yi, J. F. Li, D. Y. Wu, and Z. Q. Tian, “From plasmon-enhanced molecular spectroscopy to plasmon-mediated chemical reactions,” Nat. Rev. Chem. 2(9), 216–230 (2018). [CrossRef]  

5. Z. G. Dai, X. H. Xiao, W. Wu, Y. P. Zhang, L. Liao, S. S. Guo, J. J. Ying, C. X. Shan, M. T. Sun, and C. Z. Jiang, “Plasmon-driven reaction controlled by the number of graphene layers and localized surface plasmon distribution during optical excitation,” Light: Sci. Appl. 4(10), e342 (2015). [CrossRef]  

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

7. C. Yuan, R. Y. Liu, S. H. Wang, G. M. Han, M. Y. Han, C. L. Jiang, and Z. P. Zhang, “Single clusters of self-assembled silver nanoparticles for surface-enhanced Raman scattering sensing of a dithiocarbamate fungicide,” J. Mater. Chem. 21(40), 16264–16270 (2011). [CrossRef]  

8. Z. G. Dai, F. Mei, X. H. Xiao, L. Liao, L. Fu, J. Wang, W. Wu, S. S. Guo, X. Y. Zhao, W. Li, F. Ren, and C. Z. Jiang, ““Rings of saturn-like” nanoarrays with high number density of hot spots for surface-enhanced Raman scattering,”,” Appl. Phys. Lett. 105(3), 033515 (2014). [CrossRef]  

9. J. Yu, M. S. Yang, Z. Li, C. D. Liu, Y. S. Wei, C. Zhang, B. Y. Man, and F. C. Lei, “Hierarchical Particle-In-Quasicavity Architecture for Ultratrace In Situ Raman Sensing and Its Application in Real-Time Monitoring of Toxic Pollutants,” Anal. Chem. 92(21), 14754–14761 (2020). [CrossRef]  

10. L. P. Hou, M. R. Shao, Z. X. Li, X. F. Zhao, A. H. Liu, C. Zhang, X. W. Xiu, J. Yu, 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]  

11. Y. Quan, J. Yao, S. Yang, L. Chen, J. Li, Y. Liu, J. Lang, H. Shen, Y. Wang, Y. Wang, J. Yang, and M. Gao, “ZnO nanoparticles on MoS2 microflowers for ultrasensitive SERS detection of bisphenol A,” Mikrochim. Acta 186(8), 593 (2019). [CrossRef]  

12. R. F. Du, W. C. Yi, W. T. Li, H. F. Yang, H. Bai, J. F. Li, and G. C. Xi, “Quasi-metal Microwave Route to MoN and Mo2C Ultrafine Nanocrystalline Hollow Spheres as Surface-Enhanced Raman Scattering Substrates,” ACS Nano 14(10), 13718–13726 (2020). [CrossRef]  

13. C. D. Wang, X. H. Xu, G. Y. Qiu, W. C. Ye, Y. M. Li, R. A. Harris, and C. Y. Jiang, “Group-Targeting SERS Screening of Total Benzodiazepines Based on Large-Size (111) Faceted Silver Nanosheets Decorated with Zinc Oxide Nanoparticles,” Anal. Chem. 93(7), 3403–3410 (2021). [CrossRef]  

14. X. F. Zhao, J. Yu, Z. J. Zhang, C. H. Li, Z. Li, S. Z. Jiang, J. Pan, A. H. Liu, C. Zhang, and B. Y. Man, “Heterogeneous and cross-distributed metal structure hybridized with MoS2 as high-performance flexible SERS substrate,” Opt. Express 26(18), 23831–23843 (2018). [CrossRef]  

15. S. Tongay, H. Sahin, C. Ko, A. Luce, W. Fan, K. Liu, J. Zhou, Y. S. Huang, C. H. Ho, J. Yan, D. F. Ogletree, S. Aloni, J. Ji, S. Li, J. Li, F. M. Peeters, and J. Wu, “Monolayer behaviour in bulk ReS2 due to electronic and vibrational decoupling,” Nat. Commun. 5(1), 3252 (2014). [CrossRef]  

16. Y. Chen and M. Sun, “Two-dimensional WS2/MoS2 heterostructures: properties and applications,” Nanoscale 13(11), 5594–5619 (2021). [CrossRef]  

17. M. Rahman, K. Davey, and S. Z. Qiao, “Advent of 2D Rhenium Disulfide (ReS2): Fundamentals to Applications,” Adv. Funct. Mater. 27(10), 1606129 (2017). [CrossRef]  

18. X. J. Xu, L. H. Li, M. M. Yang, Q. L. Guo, Y. Wang, X. L. Li, X. J. Zhuang, and B. L. Liang, “Localized state effect and exciton dynamics for monolayer WS2,” Opt. Express 29(4), 5856–5866 (2021). [CrossRef]  

19. Z. X. Zou, J. W. Liang, X. H. Zhang, C. Ma, P. Xu, X. Yang, Z. X. S. Zeng, X. X. Sun, C. G. Zhu, D. L. Liang, X. J. Zhuang, D. Li, and A. L. Pan, “Liquid-Metal-Assisted Growth of Vertical GaSe/MoS2 p-n Heterojunctions for Sensitive Self-Driven Photodetectors,” ACS Nano 15(6), 10039–10047 (2021). [CrossRef]  

20. D. A. Chenet, O. B. Aslan, P. Y. Huang, C. Fan, A. M. van der Zande, T. F. Heinz, and J. C. Hone, “In-Plane Anisotropy in Mono- and Few-Layer ReS2 Probed by Raman Spectroscopy and Scanning Transmission Electron Microscopy,” Nano Lett. 15(9), 5667–5672 (2015). [CrossRef]  

21. F. Cui, C. Wang, X. Li, G. Wang, K. Liu, Z. Yang, Q. Feng, X. Liang, Z. Zhang, S. Liu, Z. Lei, Z. Liu, H. Xu, and J. Zhang, “Tellurium-Assisted Epitaxial Growth of Large-Area, Highly Crystalline ReS2 Atomic Layers on Mica Substrate,” Adv. Mater. 28(25), 5019–5024 (2016). [CrossRef]  

22. Y. Kim, B. Kang, Y. Choi, J. H. Cho, and C. Lee, “Direct synthesis of large-area continuous ReS2 films on a flexible glass at low temperature,” 2D Mater. 4(2), 025057 (2017). [CrossRef]  

23. X. F. Wang, K. Shinokita, H. E. Lim, N. B. Mohamed, Y. Miyauchi, N. T. Cuong, S. Okada, and K. Matsuda, “Direct and Indirect Exciton Dynamics in Few-Layered ReS2 Revealed by Photoluminescence and Pump-Probe Spectroscopy,” Adv. Funct. Mater. 29(6), 1806169 (2019). [CrossRef]  

24. P. Miao, J. K. Qin, Y. F. Shen, H. M. Su, J. F. Dai, B. Song, Y. C. Du, M. T. Sun, W. Zhang, H. L. Wang, C. Y. Xu, and P. Xu, “Unraveling the Raman Enhancement Mechanism on 1 T ‘-Phase ReS2 Nanosheets,” Small 14(14), 1704079 (2018). [CrossRef]  

25. J. Seo, J. Lee, Y. Kim, D. Koo, G. Lee, and H. Park, “Ultrasensitive Plasmon-Free Surface-Enhanced Raman Spectroscopy with Femtomolar Detection Limit from 2D van der Waals Heterostructure,” Nano Lett. 20(3), 1620–1630 (2020). [CrossRef]  

26. Z. Zuo, S. Zhang, Y. Wang, Y. Guo, L. Sun, K. Li, and G. Cui, “Effective plasmon coupling in conical cavities for sensitive surface enhanced Raman scattering with quantitative analysis ability,” Nanoscale 11(38), 17913–17919 (2019). [CrossRef]  

27. X. F. Zhao, C. D. Liu, J. Yu, Z. Li, L. Liu, C. H. Li, S. C. Xu, W. F. Li, B. Y. 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]  

28. X. G. Zhang, X. L. Zhang, C. L. Luo, Z. Q. Liu, Y. Y. Chen, S. L. Dong, C. Z. Jiang, S. K. Yang, F. B. Wang, and X. H. Xiao, “Volume-Enhanced Raman Scattering Detection of Viruses,” Small 15(11), 1805516 (2019). [CrossRef]  

29. S. S. Singha, S. Mondal, T. S. Bhattacharya, L. Das, K. Sen, B. Satpati, K. Das, and A. Singha, “Au nanoparticles functionalized 3D-MoS2 nanoflower: An efficient SERS matrix for biomolecule sensing,” Biosens. Bioelectron. 119, 10–17 (2018). [CrossRef]  

30. R. Yang, Y. Q. Cheng, Y. J. Song, V. I. Belotelov, and M. T. Sun, “Plasmon and Plexciton Driven Interfacial Catalytic Reactions,” Chem. Rec. 21(4), 797–819 (2021). [CrossRef]  

31. K.H. Lee, H. Jang, Y.S. Kim, C.H. Lee, S.H. Cho, M. Kim, H. Son, K.B. Bae, D.V. Dao, Y.S. Jung, and I.H. Lee, “Synergistic SERS Enhancement in GaN-Ag Hybrid System toward Label-Free and Multiplexed Detection of Antibiotics in Aqueous Solutions,” Adv. Sci.e2100640 (2021), early access. [CrossRef]  

32. B. Han, L. Chen, S. Jin, S. Guo, J. Park, H. S. Yoo, J. H. Park, B. Zhao, and Y. M. Jung, “Modulating Mechanism of the LSPR and SERS in Ag/ITO Film: Carrier Density Effect,” J. Phys. Chem. Lett. 12(31), 7612–7618 (2021). [CrossRef]  

33. L. Tao, Z. Y. Li, K. Chen, Y. Q. Zhou, H. Li, X. M. Wang, R. Z. Zhan, X. Y. Hou, Y. Zhao, J. L. Xu, T. Qiu, X. Wan, and J. B. Xu, “A spontaneously formed plasmonic-MoTe2 hybrid platform for ultrasensitive Raman enhancement,” Cell Rep. Phys. Sci. 2, 100526 (2021). [CrossRef]  

34. S. Jiang, L. Zhao, Y. Shi, C. Xie, N. Zhang, Z. Zhang, Y. Huan, P. Yang, M. Hong, X. Zhou, J. Shi, Q. Zhang, and Y. Zhang, “Temperature-dependent Raman spectroscopy studies of the interface coupling effect of monolayer ReSe2 single crystals on Au foils,” Nanotechnology 29(20), 204003 (2018). [CrossRef]  

35. Z. Li, S. Z. Jiang, Y. Y. Huo, A. H. Liu, C. Zhang, J. Yu, M. H. Wang, C. H. Li, Z. Y. Lu, and B. Y. Man, “3D Hybrid Plasmonic Nanostructures with Dense Hot Spots Using Monolayer MoS2 as Sub-Nanometer Spacer,” Adv. Mater. Interfaces 5(19), 1800661 (2018). [CrossRef]  

36. W. Y. Zhang, P. H. Man, M. H. Wang, Y. Shi, Y. Y. Xu, Z. Li, C. Yang, and B. Y. Man, “Roles of graphene nanogap for the AgNFs electrodeposition on the woven Cu net as flexible substrate and its application in SERS,” Carbon 133, 300–305 (2018). [CrossRef]  

37. M. Z. Li, Y. J. Wei, C. Fan, G. Q. Li, Q. Hao, and T. Qiu, “Mixed-dimensional van der Waals heterojunction-enhanced Raman scattering,” Nano Res. (2021), https://doi.org/10.1007/s12274-021-3537-2.

38. M. P. Chen, B. Ji, Z. Y. Dai, X. Y. Du, B. C. He, G. Chen, D. Liu, S. Chen, K. H. Lo, S. P. Wang, B. P. Zhou, and H. Pan, “Vertically-aligned 1 T/2H-MS2 (M = Mo, W) nanosheets for surface-enhanced Raman scattering with long-term stability and large-scale uniformity,” Appl. Surf. Sci. 527, 146769 (2020). [CrossRef]  

39. Y. R. Fang, Y. Z. Li, H. X. Xu, and M. T. Sun, “Ascertaining p,p ‘-Dimercaptoazobenzene Produced from p-Aminothiophenol by Selective Catalytic Coupling Reaction on Silver Nanoparticles,” Langmuir 26(11), 7737–7746 (2010). [CrossRef]  

40. B. Dong, Y. R. Fang, X. W. Chen, H. X. Xu, and M. T. Sun, “Substrate-, Wavelength-, and Time-Dependent Plasmon-Assisted Surface Catalysis Reaction of 4-Nitrobenzenethiol Dimerizing to p,p ‘-Dimercaptoazobenzene on Au, Ag, and Cu Films,” Langmuir 27(17), 10677–10682 (2011). [CrossRef]  

41. W. H. Lin, Y. Q. Cao, P. J. Wang, and M. T. Sun, “Unified Treatment for Plasmon Exciton Co-driven Reduction and Oxidation Reactions,” Langmuir 33(43), 12102–12107 (2017). [CrossRef]  

42. M. T. Sun and H. X. Xu, “A Novel Application of Plasmonics: Plasmon-Driven Surface-Catalyzed Reactions,” Small 8(18), 2777–2786 (2012). [CrossRef]  

43. Z. L. Zhang, P. Xu, X. Z. Yang, W. J. Liang, and M. T. Sun, “Surface plasmon-driven photocatalysis in ambient, aqueous and high-vacuum monitored by SERS and TERS,” J. Photochem. Photobiol., C 27, 100–112 (2016). [CrossRef]  

44. C. H. Li, S. C. Xu, J. Yu, Z. Li, W. M. 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]  

45. 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,” Nanophotonics 10(5), 1529–1540 (2021). [CrossRef]  

46. Q. Zhang, W. J. Wang, J. Q. Zhang, X. H. Zhu, and L. Fu, “Thermally Induced Bending of ReS2 Nanowalls,” Adv. Mater. 30(3), 1704585 (2018). [CrossRef]  

Supplementary Material (1)

NameDescription
Supplement 1       Supplement 1: Supplementary data results

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 (9)

Fig. 1.
Fig. 1. Schematic diagram of the fabrication of PSi/Au/ReS2NC/Ag.
Fig. 2.
Fig. 2. (A) SEM image of PSi/Au/ReS2NC. (B) SEM image of PSi/Au/ReS2NC/Ag. (C) and (D) HRTEM image of PSi/Au/ReS2NC/Ag. (E) EDS elemental mappings from Au, Re, S, Ag and Si on the PSi/Au/ReS2NC/Ag.
Fig. 3.
Fig. 3. (A) XPS spectra of Re of PSi/Au/ReS2NC and PSi/Au/ReS2NC/Ag sample. (B) XPS spectra of Ag of PSi/Au/ReS2NC/Ag sample. (C) The Raman spectra of PSi/Au/ReS2NC and PSi/Au/ReS2NC/Ag sample. (D) The scanning Raman mapping of E2g mode of the PSi/Au/ReS2NC sample.
Fig. 4.
Fig. 4. (A) SERS spectrum of R6G (10−5 M) collected on the PSi/Au/ReS2NC/Ag with varying deposition thickness of Ag. (B) The broken line graphs of SERS intensities at 614 and 776 cm-1 with Ag deposition thickness. (C)-(G) Simulated vertical electric field distributions of the ReS2NC deposited different Ag thickness of (C) 5 nm, (D) 10 nm, (E) 15 nm, (F) 20 nm, (G) 25 nm. (H) The simulated enhancement factor (|E|2/|E0|2) of the different hybrid structures.
Fig. 5.
Fig. 5. Schematic diagram of the proposed charge transfer mechanism.
Fig. 6.
Fig. 6. (A)-(F) In-situ SERS monitoring the dimerization of PNTP to DMAB for PSi/Au/ReS2NC/Ag with different deposited thickness of Ag. (G) Time dependence of ${I_{{V_{({N = N} )}}}}/{I_{{V_{({N{O_2}} )}}}}$ for PSi/Au/ReS2NC/Ag with different Ag thickness under 532 nm laser excitation. (H) Ag thickness-dependent SERS intensities of the ${V_{({N = N} )\; }}$ band.
Fig. 7.
Fig. 7. (A) SERS spectrum of R6G (10−5 - 10−11 M) in PSi/Au/ReS2NC/Ag. (B) The relation between the intensity of SERS signals and the concentration of R6G at 614 and 776 cm−1 in logarithmic coordinates. (C) The scanning Raman mappings of characteristic peak of R6G (10−6 M) at 614 cm-1. (D) SERS spectrum of R6G (10−6 M) collected on different substrates. SERS intensity at 614 cm-1 (E) and 776 cm-1 (F) obtained from Raman spectra in (D).
Fig. 8.
Fig. 8. (A) SERS spectrum of R6G (10−5) in PSi/Au/ReS2NC at 25 - 85 °C. (B) The broken line graphs of SERS signal intensity at 614 and 776 cm-1 with temperature in PSi/Au/ReS2NC. (C) SERS spectrum of R6G (10−5) in PSi/Au/ReS2NC/Ag at 25 - 85 °C. (D) The broken line graphs of SERS signal intensity at 614 and 776 cm-1 with temperature in PSi/Au/ReS2NC/Ag. (E) Schematic diagram of the morphology of ReS2 nanosheets changing with the temperature.
Fig. 9.
Fig. 9. (A) SERS spectrum of MG (10−4 - 10−9 M) in lake water solution collected on PSi/Au/ReS2NC/Ag. (B) The relation between the intensity of SERS signals and the concentration of MG at 1620 cm−1 in logarithmic coordinates. (C) SERS spectrum of CV (10−6 - 10−10 M) in lake water solution collected on PSi/Au/ReS2NC/Ag. (D) The relation between the intensity of SERS signals and the concentration of CV at 915 cm−1 in logarithmic coordinates.

Equations (1)

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

R S D = i = 1 n ( I i I ¯ ) 2 n 1 I ¯ ,
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.