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

Colorimetric and SERS dual-mode detection of lead Ions based on Au-Ag core-shell nanospheres: featuring quick screening with ultra-high sensitivity

Open Access Open Access

Abstract

A novel strategy for colorimetric and surface-enhanced Raman scattering (SERS) dual-mode sensing of the lead ions (Pb2+) was established based on gluconate ion (Gluc) modified and 2-Naphthalenethiol (2-NT) tagged Au-Ag core-shell nanoparticles (NPs). Due to the complex formation between adsorbed Gluc and Pb2+, the addition of Pb2+ can induce the aggregation of Gluc/2-NT@Au@Ag NPs. Correspondingly, the aggregated Gluc/2-NT@Au@Ag NPs caused a significant difference in the color and SERS intensity. As a result, such Gluc/2-NT@Au@Ag NPs can achieve the sensing of Pb2+ using both colorimetric and SERS signals as the indicator, which features with wide response range from 10−11 to 10−5 M, rapid screening and high sensitivity (with a limit of detection (LOD) of 0.185 pM). Furthermore, such dual-mode sensor was demonstrated not to be responsive to other cations, and facilitate the sensing of real samples in practical environment. With rapid screening ability and outstanding sensitivity, we anticipate that this method would holding great potential for the applications in environmental monitoring.

© 2019 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

With the high level of industrial development, lead and its compounds have been applied to various aspects including lead-acid batteries, industrial plating and smelting production [1]. But when enjoying the convenience brought by high-level industry, we have to suffer from the troubles of heavy metal contamination. For instance, waste water from the lead smelting production contains Pb2+. With improper treatment, Pb2+ will be enriched in water or land eventually and cannot be degraded, which is seriously detrimental to the environment. For humans, Pb2+ can cause varieties of adverse effects even at a very low level [2]. Human may suffer from anemia, neurological dysfunction, reproductive system damage and kidney damage induced by Pb2+ [3,4]. Blood lead levels, even as low as 50 µg/L, can still damage the baby's brain and cause lifelong health problems [5]. Therefore, Pb2+ has been listed as a notorious toxic environmental pollutant. Drinking water is a major source of lead exposure for human, and the world health organization (WHO) has set a standard of 10 µg/L for Pb2+ in drinking water [6]. Up to now, various Pb2+ trace detection techniques have been developed, such as atomic absorption spectrometry (AAS) [79], inductively coupled plasma mass spectrometry (ICP-MS) [10,11], inductively coupled plasma-atomic emission spectroscopy (ICP-AES) [12] and fluorescence spectroscopy [1318]. Although these traditional methods exhibit excellent sensitivity, the requirement of cumbersome steps and the extended times limit their wide applications in Pb2+ detection on site. Thus, there is a great demand of new strategies for easy and rapid detection of Pb2+.

Colorimetric sensing approaches of heavy metal ions has been of interests because they do not depend on the advanced instruments and often response quickly. To date, Memon et al. reported that the truncated DNAzyme was split into single-stranded DNA fragments in the presence of Pb2+. The detached DNA fragments were then adsorbed onto gold NPs and protected against salt-induced aggregation, which would result in rapid color change from blue to pink [19]. Rong et al. designed a Pb2+ direct detection platform based on magnetic nano-enzyme, magnetic beads and gold nanoparticles, which can be used for colorimetric detection of lead ions [20]. Sang et al. presented a fast, simple and sensitive colorimetric sensor that can use tyrosine-functionalized gold NPs to simultaneously (or separately) detect Cr3+ and Pb2+ [21]. Guo et al. developed cost-effective and rapid colorimetric method for simultaneous detection of Hg2+, Pb2+and Cu2+ using papain enzyme-functionalized gold NPs with a detection limit of 0.2 µM [22]. However, colorimetric analysis may be prone to interference by the colors of samples. Moreover, the detection limit of some colorimetric methods is sometimes not satisfactory.

SERS which can amplify the Raman signals by several orders of magnitude (as high as 1015) using plasmonic nanostructures, shows to be another choice for highly selective and sensitive detection. For example, Olga Guselnikova et al. designed a surface plasmon-polariton based platform to realized selective and sensitive detection of heavy metal ions (with LOD of 10−14 M) [23]. They also developed a dual-mode functional chip for stereoselective discrimination of chiral amines using Wettability-Based mobile and portable SERS measurements [24]. Mark S. Frost et al. utilized a SERS-active citrate-functionalized gold sensor to detect Pb2+ through the metal affinity of citrate molecules [25]; Fu et al. proposed a highly sensitive and selective SERS method for determining Pb2+ based on a DNAzyme-linked plasmonic nanomachine and this method realized the lowest detection concentration of Pb2+ to 1.0 nM [26]; Shi et al. reported a kind of poly adenosine assisted SERS silicon wafer composed of core(Ag)-satellite(Au) NPs modified silicon wafer, which was used for high-performance Pb2+ detection [27]; Zou et al. used oligonucleotide functionalization and gold-plated polystyrene microspheres to construct a new conceptual DNA logic gate for synchronous detection of Hg2+ and Pb2+ [28]. Despite the impressive progress has been achieved, the applicability and sensitivity of Pb2+ sensing approaches still need improvement especially in the practical samples.

Taking the above-mentioned into account, herein, we present a dual-mode sensing method based on colorimetry and SERS for the detection of Pb2+. The principle of this method is illustrated in Fig. 1. Generally, gold NPs are first coated with silver nanoshells by the reduction of silver nitrate using glucose. Compared to Au or Ag NPs, Au@Ag core/shell structures have more advantages which combine the strong SERS activity of Ag and well stability of Au. In the growth of the silver shells, Gluc is adsorbed onto the surface of silver shells as the stabilizing agent [2931] . The Gluc stabilized Au @ Ag NPs are then tagged with Raman reporter 2-NT through the Au-S bonds, which results in the formation of Gluc/2-NT@Au@Ag NPs. In such core-shell nanostructure, Au@Ag NPs are employed as SERS generators to boost the Raman signals of 2-NT. And the adsorbed Gluc can selectively form a complex with the lead ion due to the affinity between carboxyl and hydroxyl groups in Gluc and Pb2+ [32]. Upon the addition of Pb2+, the nanoprobes are aggregated and the hot spots are simultaneously formed, which consequently induced the obvious color change and significant enhancement of the Raman signals of 2-NT. This method combines the advantages of both colorimetry and SERS techniques, covering a broad response range with easy operation, fast readout and excellent sensitivity. This is particularly favorable with a large number of samples, as we can implement rapid preliminary screening with naked eyes and subsequently employ SERS spectra for quantitative analysis.

 figure: Fig. 1.

Fig. 1. (A) The synthesis principle of Gluc/2-NT@Au@Ag NPs. (B) Schematic illustration of SERS and colorimetric determination for Pb2+.

Download Full Size | PDF

2. Experimental section

2.1 Chemicals and reagents

Chloroauric acid tetrahydrate (HAuCl4·4H2O), sodium citrate dihydrate (C6H5Na3O7•2H2O ≥ 99.0%), silver nitrate (AgNO3), sodium hydroxide (NaOH), D (+)-Glucose (C6H12O6) and D-Gluconic acid sodium salt (C6H11NaO7, ≥99.0%) were purchased from MACKLIN (Shanghai, China). The chloride salts of Ni2+, Cu2+, Cr2+, Ca2+, Zn2+, K+, Hg2+, Na+ and Pb2+ were purchased from MACKLIN (Shanghai, China). 2-Naphthalenethiol (C10H7SH, ≥99.0%) were purchased from Sigma-Aldrich (Shanghai, China). All the reagents were used as received and deionized water with a resistivity of 18.25 M Ω · cm was used in all of the experiments. The real samples of tap water and drinking water were collected from Shanghai, China.

2.2 Instruments

Extinction spectra were recorded by a Shimadzu UV-3600 PC spectrophotometer with quartz cuvettes of 1 cm path length. Transmission electron microscopy (TEM) images were obtained with an JEM-2100F, JEOL electron microscope operating at 200 kV. SERS spectra were collected by a confocal Raman spectrometer (XploRA Plus, HORIBA) at the excitation of 638 nm. The scattering light was collected by a 10X objective lens to CCD. All the SERS spectra were acquired with two accumulations and for each accumulation, the exposure time was 4 seconds.

2.3 Preparation of Gluc/2-NT@Au@Ag NPs

According to the method of producing gold seeds with G. Frens [33], gold NPs were synthesized by reducing chloroauric acid with sodium citrate. Typically, 306 µL of HAuCl4·4H2O (10% w/w) was dissolved in 99.694 mL of deionized water. The mixture was heated to boiling. Then, 10 mL of 47.9 mM sodium citrate dihydrate was added into the boiled mixture. The mixed solution was kept on boiling for 15 min and then allowed to cool to room temperature.

6 mL of the resulting gold NPs were centrifuged at 9500 r/min for 35 min to remove excessive sodium citrate. The obtained gold NPs were dissolved in 54 mL of deionized water, and the mixture was then heated to 60°C under rapid agitation. Subsequently, 500 µL of 0.08 M D (+)-Glucose solution and 140 µL of 0.1 M AgNO3 solution were added, followed by the addition of NaOH solution (1 M, 200 µL). After an interval of 3 minutes, 5 mL of 37.8 mM D- Gluconic acid sodium salt was added to the mixed solution at a rate of 200 µL every 10 seconds. The mixture was further heated for 30 minutes. After the color of the resulting solution became orange, the solution was cooled to room temperature and kept in the dark for at least one week. During this period, Gluc containing hydroxyl groups and carboxyl group could be bounded to the silver shell by physical adsorption more efficiently [34].

6 mL of the above Gluc/2-NT@Au@Ag NPs solution was added by 30 µL of 2-NT (100 mM). The mixture was oscillated for 2 h and kept in dark for later use.

2.4 Colorimetric and SERS detection of Pb2+

First of all, the detection of Pb2+ was conducted by colorimetric method. Specifically, 20 µL of aqueous solutions containing Pb2+ with different concentrations (ranging from 10−4 to 10−6 M) were added to 180 µL of the Gluc/2-NT@Au@Ag solution. The mixtures were then photographed and subjected to extinction detection by UV-3600 PC spectrophotometer. In the experiments, the color of the mixture with higher concentration of Pb2+ changed faster and more remarkably. Correspondingly, the recorded extinction spectra with an obvious change could be used for quantitative analysis.

Despite the colorimetric sensing is fast and efficient, the relatively low detection limit is not suitable for trace sensing. For trace sensing of the lead, SERS-based detection shows an advantage. In a specific experiment, aqueous solutions containing Pb2+ (10−6–10−11 M, 20 µL) were added to 180 µL of the Gluc/2-NT@Au@Ag NPs solution. After 45 minutes, Raman spectra of the mixtures were recorded. Apparently, as the concentration of Pb2+ increased, the Raman signals of 2-NT were dramatically enhanced. The SERS spectra were collected for the quantitative sensing of Pb2+.

2.5 Selective detection of Pb2+

To test the selectivity of this presented method for the detection of Pb2+, several ubiquitous metal ions Ni2+, Cu2+, Cr2+, Zn2+, Ca2+, K+, Ag+, Hg2+ and Na+ aqueous solutions were used as control. In this experiment, the solutions of various metal ions (10−5 M, 20 µL) were added into 180 µL of the Gluc/2-NT@Au@Ag NPs solution. Only in the presence of Pb2+, the color and the extinction spectra of the solution was clearly changed. As for other metal ions, no obvious change was observed whether for the color or extinction spectra.

2.6 The detection of the real samples

To confirm the feasibility of the nanosensor for the application in real sample such as tap water and drinking water, the recovery experiments were performed under present conditions by adding Pb2+ spiked real samples into the Gluc/2-NT@Au@Ag solutions. The LSPR band at 417 nm and Raman band at 1621 cm−1 of all the solutions were used to calculate Pb2+ concentration in the samples according to the calibration graphs, which could be treated as recovery results, and then the recovery rate was calculated by the formula (see Appendix A for details).

2.7 Statistical Analysis

Actually, for colorimetry sensing, 10 samples with Pb2+(10−6–10−4 M) were tested and the given spectrum for each sample was averaged from 15 measurements. While for SERS sensing, 6 samples with Pb2+(10−11–10−6 M) were tested and the given spectrum for each sample was averaged from 15 measurements. All the experiments were repeated 5 time. Similarly, the blank sample and other metal ions samples were tested in the same way.

3. Results and discussion

3.1 Characterization of Gluc/2-NT@Au@Ag NPs

The preparation of Gluc/2-NT@Au@Ag NPs contained three steps. First, the gold NPs were synthesized through the reduction of HAuCl4·4H2O mediated by sodium citrate dihydrate. Au NPs were examined by TEM and UV-vis spectroscopy. Figure 2A showed that the gold NPs were spherical and monodispersed with a size about 12 nm. As indicated in Fig. 3A(a), an obvious absorption band at 520 nm appeared, which was typical for the surface plasmon resonance of gold NPs. And the gold colloid solution was burgundy. Then, Au NPs were coated by silver shells through the reduction of AgNO3 by D (+)-Glucose, and simultaneously Gluc was acted as capping agent. After the silver shell coating, the color changed from burgundy to yellow. With the increase of the thickness of silver shell, the original absorption peak at 520 nm was blue-shifted to 411 nm (Fig. 3A(b)) due to the increased particle volume and the dielectric properties of silver [35]. Thus, Gluc capped Au-Ag core-shell nanostructures were synthesized (denoted as Gluc@Au@Ag). Gluc@Au@Ag NPs were then observed by TEM which are monodispersed with an averaged diameter of about 22 nm (Fig. 2B). Finally, the Raman reporter 2-NT was tagged onto the surface of Gluc@Au@Ag via a Ag-S bond. The obtained 2-NT tagged Gluc@Au@Ag was also observed by TEM (Fig. 2C) which showed a similar morphology as Gluc@Au@Ag NPs. Moreover, the SPR band of Gluc/2-NT@Au@Ag NPs red-shifts to 417nm (Fig. 3A(c)). The obtained Gluc/2-NT@Au@Ag NPs are expected to keep their colloidal stability before the utilization for sensing. As shown in Fig. 2(E) and 2(F), Gluc/2-NT@Au@Ag NPs preserved the hydrodynamic diameter and absorption band over time, which indicated that Gluc/2-NT@Au@Ag NPs were quite stable. SERS performance of Gluc/2-NT@Au@Ag NPs were then investigated, which proved such Gluc/2-NT@Au@Ag NPs as homogeneous substrates (Fig. 2(G) and 2(H)). Gluc/2-NT@Au@Ag NPs exhibited characteristic SERS bands of 2-NT at 422, 842, 1064, 1379 and 1621cm−1 [33] (see Appendix B for detailed assignments). We thereafter conducted a preliminary experiment. While 7.5E-6 M Pb2+ was added into Gluc/2-NT@Au@Ag NPs, the NPs were found to aggregate quickly accompanied by a significant change of the color (Fig. 3A). This is reasonable because the Gluc on the silver shells can bind with Pb2+ to form a complex. Simultaneously, as shown in Fig. 3B, the SERS signals of Gluc/2-NT@Au@Ag in the presence of Pb2+ were observed to be intensely amplified which was caused by the hotspots occurred in the aggregates. The underlying mechanism of the hot-spots based Raman enhancing were then demonstrated by performing the simulation of electromagnetic field distributions in the Au@Ag NPs aggregates with FDTD solutions. According to TEM images in Fig. 2D, in the presence of the lead ions, the distances between Au@Ag NPs were counted to be around 1.72 nm. Then different possible the aggregation states of Au@Ag NPs were simulated at the interdistance of 1.72 nm(see Appendix C for the simulation of electromagnetic field distributions). Compared with one single metal particle, the aggregates exhibited the stronger electromagnetic coupling effect due to the inherent interparticle hot-spots. And as more particles formed the aggregates, more hot-spots were created which correspondingly amplify the Raman vibrations. The FDTD calculation is in agreement with experimental measurements, which provide some certain theoretical support for experiments. Such preliminary results facilitated the feasibility of using Gluc/2-NT@Au@Ag for colorimetric and SERS dual-mode sensing of the lead ions.

 figure: Fig. 2.

Fig. 2. TEM images of (A) Au NPs. (B) Gluc@Au@Ag NPs. (C) Gluc/2-NT@Au@Ag NPs. (D) The aggregations of Gluc/2-NT@Au@Ag NPs in the presence Pb2+ of 7.5E-6 M. (E) The extinction spectra of the Gluc/2-NT@Au@Ag NPs suspension (without the addition of Pb2+) for 5 days. (F) The hydrodynamic size of the Gluc/2-NT@Au@Ag NPs suspension (without the addition of Pb2+) for 5 days. (G)The histogram of SERS intensity of Gluc/2-NT-@Au@Ag NPs at 1621 cm−1 band. (inset: SERS mapping image integrated at 1621 cm−1 band and the corresponding bright-field image). (H) A serials of SERS spectra for all mapping blocks in (G) (RSD of the intensities of 1621 cm−1 band is calculated to be 9.365%).

Download Full Size | PDF

 figure: Fig. 3.

Fig. 3. (A) Extinction spectra of (a) Au NPs, (b) Gluc@Au@Ag NPs, (c) Gluc/2-NT@Au@Ag NPs and (d) Gluc/2-NT@Au@Ag NPs aggregations in the presence Pb2+ of 7.5E-6 M. (inset: Corresponding photographs (Sequencing from left to right: Au NPs, Gluc@Au@Ag NPs, Gluc/2-NT@Au@Ag NPs, the aggregations of Gluc/2-NT@Au@Ag NPs in the presence Pb2+)). (B) The SERS spectra of (a) Gluc/2-NT@Au@Ag NPs aggregations in the presence Pb2+ of 7.5E-6 M. (b) Gluc/2-NT@Au@Ag NPs and (c) Gluc@Au@Ag NPs.

Download Full Size | PDF

3.2 Detection of Pb2+ based on Gluc/2-NT@Au@Ag

3.2.1 Colorimetric sensing of Pb2+

To analyze the quantitative relationship between various Pb2+ concentrations and corresponding extinction spectra intensity, the as-prepared nanoprobe (Gluc/2-NT@Au@Ag) were used to achieve the detection of Pb2+. First, time-sensitive extinction spectra of the dispersion were recorded after the addition of Pb2+. As shown in Fig. 4(A) and 4(B), Gluc/2-NT@Au@Ag NPs suffer from different aggregation stages. The relatively stable stage was between 30 min to 120 min. In our experiments, the extinction spectra at 45 min were used. Then, different concentrations of Pb2+ were added into the dispersion of the Gluc/2-NT@Au@Ag NPs. The response of Gluc/2-NT@Au@Ag was recorded. As shown in the Fig. 4D, after the addition of Pb2+ with concentrations ranging from 10−6 to 10−4 M, the colors of the dispersions of Gluc/2-NT@Au@Ag changed rapidly from yellow to orange and finally dark green, which were easily distinguishable by naked eyes. This is because the addition of Pb2+ induces the aggregation of Gluc/2-NT@Au@Ag NPs through the affinity between them. The extinction spectra of the dispersions were investigated after 45 min which further proved the occurrence of aggregation. It was noted that the SPR band of Gluc/2-NT@Au@Ag at 417 nm gradually decreased as the concentration of Pb2+ increased and a new broad band around 570–640 nm appeared in the extinction spectra (Fig. 4C). Herein, the linear dependence between the intensity of the absorption band at 417 nm and Pb2+ concentration was obtained. The calibration curve in the Pb2+ concentration ranging from 2.5E-5 to 1E-6 M was plotted in Fig. 4D. The relationship between the decrease in the SPR band intensity (Y) of Gluc/2-NT@Au@Ag at 417 nm and Pb2+ concentration (X) was regressed linearly as the following equation:

$$Y = 0.31786 - 7513.53449\textrm{X}$$
It could be seen that the calibration curve exhibited a good linear relationship with the concentration from 2.5E-5 to 1E-6 M and its R2 value was 0.96527 (see Table 1 for details). Based on these results, the LOD could be calculated to be 0.252 µM with the 3σ/s method. In the method, the symbols ‘σ’ and ‘s’ were denoted as standard deviation (SD) (see the Appendix D for the formulas) of the blank signal and the slope of calibration curve, respectively [34]. Therefore, in field analysis of mass samples, those ones with concentrations higher than 10−5 M could be quickly screened out by directly observing the colors, and the concentration in the linear range could be calculated.

 figure: Fig. 4.

Fig. 4. (A) Extinction spectra of Gluc/2-NT@Au@Ag NPs in presence of 1E-5 M Pb2+ over time. (B) Plot of absorbance of SPR band at 417 nm with time in presence of different concentration of Pb2+ ions (1E-6 M, 5E-6 M and 1E-5 M). (C) Extinction spectral change of Gluc@Au@Ag NPs in presence of Pb2+ (10−4–10−6 M). (D) Linear plot of absorbance of SPR band at 417 nm at different concentrations of Pb2+. The error bars indicate the standard deviation of 15 measurements.

Download Full Size | PDF

Tables Icon

Table 1. Linear plot of absorption band at 417 nm and SERS intensity of the 1621 cm−1 band upon the addition of Pb2+ in different concentrations.

When the Pb2+ concentration was down to 10−6 M, no obvious color change of the nanoprobe could be observed with the addition of Pb2+. Similarly, there is no apparent change in the SPR band of the nanoprobe after the addition of Pb2+ at a level of 10−6 M. Hence, a more sensitive sensing method is necessary for lower concentration detection of Pb2+.

3.2.2 SERS sensing of Pb2+

As illustrated before, the addition of Pb2+ can induce the aggregation of the nanoprobes. As the concentration of Pb2+ decreased, the lower aggregation degree was obtained corresponding with weaker SERS signals. To achieve the sensing of Pb2+, the SERS signals were estimated. The lead ions with a concentration range from 10−6 to 10−11 M were added to Gluc/2-NT@Au@Ag dispersions, and the mixtures were then subject to SERS measurements by the confocal Raman spectrometer. As shown in Fig. 5A, the characteristic Raman peaks of 2-NT at 1621 cm−1 could be clearly detected. As the Pb2+ concentrations increased, the Raman signals of the Gluc/2-NT@Au@Ag were observed to be enhanced at different levels. This is because the formation of hotspots between the Gluc/2-NT@Au@Ag NPs. And the results showed that hotpots responsive SERS signals were extremely sensitive to detect the Pb2+. In our experiments, the Raman intensity at 1621 cm−1 showed good linear relationship with the logarithm of Pb2+ concentration from 10−6 M to 10−11 M. A calibration curve was shown in Fig. 5B and it is linearly regressed as the equation:

$$Y = 8997.46\textrm{9} + 706.525 \times L\textrm{og}_{10}X$$
The R2 value was 0.9919 (see Table 1 for details) and the LOD was calculated to be 0.185 pM by triple SNR method, which was far below the international safety standard for Pb2+ (0.01 mg/L) [7] (see the Appendix D for the formulas of the specific calculation method and the LOD).

 figure: Fig. 5.

Fig. 5. (A) Raman spectra of Gluc/2-NT@Au@Ag NPs in presence of Pb2+ (10−11–10−6 M). (B) Linear plot of Raman intensity of the 1621 cm−1 peak at various concentrations of Pb2+. The error bars indicate the standard deviation of 15 measurements.

Download Full Size | PDF

By combining the colorimetric and SERS methods, the detection of Pb2+ covered a wide range from 10−11 to 10−5 M. Interestingly, for mass samples analysis on-site, those with higher concentration of Pb2+ (> 10−5 M) could be rapidly screened out by naked eyes. While for samples with lower concentration of Pb2+(10−6–10−11 M), the accurate concentration of Pb2+ in the sample could be detected using SERS method. In order to test the reliability of the presented method, all experiments were repeated 5 times using different batches and the results follow the same law. Generally, our method has a lower detection limit and is more convenient to operate (Table 2). Hence, the presented dual-mode sensing method based on colorimeter and SERS combining rapid readout and excellent sensitivity showed its superiority in the detection of Pb2+ especially in field test.

Tables Icon

Table 2. The comparison of the detection of Pb2+ with previously reported methods.

3.3 Selectivity of the nanoprobe

The selectivity of the developed dual-mode sensing of Pb2+ was further examined by testing eight prevalent metal ions (Ni2+, Cu2+, Cr2+, Ca2+, Zn2+, K+, Ag+, Hg2+, Na+). Specifically, the above mentioned metal ions (10−5 M) were added to 180 µL of Gluc/2-NT@Au@Ag, and the mixtures were incubated for 45 min. Meanwhile, Gluc/2-NT@Au@Ag treated with 10−5 M Pb2+ under the same conditions was used as a positive control and Gluc/2-NT@Au@Ag without any treatment as a negative control. All mixtures were then photographed and subject to UV-vis absorption measurement. As revealed in Fig. 6A, only the color of the mixture containing Pb2+ changed obviously, while the colors of the mixtures containing other metal ions remained substantially the same as that of the negative control. Similarly, as indicated by the extinction spectra, the SPR band decrease of the nanoprobe was only observed in the presence of Pb2+, owing to the strong affinity between the gluconate ions on surface of the nanoprobe and the lead ions (Fig. 6A). Quantitative evaluation of the selectivity for different ions based on the intensity of SPR band at 417 nm was shown in Fig. 6B, which definitely showed the presented method had an outstanding selectivity for Pb2+, and could specifically distinguish Pb2+ from interfering metal ions. Similarly, SERS intensity of Gluc/2-NT@Au@Ag NPs with addition of Pb2+ (10−6 M) is significantly increased compared to that of Gluc/2-NT@Au@Ag NPs with other metal (10−6 M) (Fig. 6C and 6D), which prove that SERS sensing approach is also capable of selective detection of Pb2+. Therefore, the dual-detection method we proposed is capable of selectively detecting Pb2+.

 figure: Fig. 6.

Fig. 6. Selectivity test with various metal ions. (A) Histogram of absorbance of SPR band at 417 nm in the presence of different metal ions. (inset: Photographs of Gluc/2-NT@Au@Ag NPs in the presence of different metal ions of 10−5 M). (B) Extinction spectra of Gluc/2-NT@Au@Ag NPs in the presence of different metal ions. (C) Histogram of Raman intensity of the 1621 cm−1 band in the presence of different metal ions. The error bars indicate the standard deviation of 15 measurements. (D) Raman intensity of the 1621 cm−1 band in the presence of different metal ions of 10−6 M.

Download Full Size | PDF

3.4 Detection of Pb2+ in real sample

Tap water and drinking water were employed to further evaluate the application of presented method in real samples. The water sample was centrifuged to remove other impurities. The recovery experiments were performed from tap-water and drinking water samples spiked with 10−5 and 10−9 M of Pb2+, respectively. Then the results of the SPR bands and the Raman signals were respectively used to calculate Pb2+ concentration in the samples from the calibration graphs. The standard concentration, average detection concentration and recovery rate of each sample was listed in Table 3. Obviously, there was no significant difference between the determination results and the standard solution. A conclusion could be drawn that the recoveries of this presented method in real samples were in an acceptable range of 129.1% - 113.1%, indicating that such dual-mode sensing method had great potential in real samples.

Tables Icon

Table 3. Determination of Pb2+ spiked in the real water sample

4. Conclusions

In summary, a new strategy for the detection of Pb2+ using Gluc modified and 2-NT tagged Au-Ag core-shell nanoprobes had been developed. Upon the addition of Pb2+, the nanoprobes were induced to aggregate, correspondingly the color and SERS signals of nanoprobes changed. This facilitated the colorimetric and SERS dual-mode sensing of Pb2+. Combining both advantages of colorimeter and SERS, the detection method covered a wide range from 10−5 to 10−11 M with a LOD of 0.185 pM. Moreover, this method had an outstanding selectivity for Pb2+. And the applicability in the real samples was also demonstrated with an excellent recovery rate. The results showed that such dual-mode method was extremely powerful for the detection of Pb2+ with a broad response range, rapid screening and high sensitivity.

Appendix A: Calculation method of recovery rate

The percentage of recovery values could be expressed by the following formula:

$$\textrm{Recovery}(\%) = [(C_d/Cs)] \times 100\%$$
Where, “Cd” is the spiked concentration of Pb2+ in the real water samples; “Cs” is the average detection concentration of Pb2+ in the real water.

Appendix B: The fundamental vibrational wavenumbers of 2-naphthalenethiol

Band assignments of major peaks for 2-naphthalenethiol based on the literatures [36] are as follows in Table 4:

Tables Icon

Table 4. Observed fundamental vibrational wavenumbers of 2-naphthalenethiol

Appendix C: The simulation of the electric field distributions of the Au@Ag NPs

In the simulation, the excitation wavelength was set to be 638 nm. The diameter of Au@Ag NPs was set to be 12 nm for Au core and 5 nm for Ag shell. Besides, the light source was perpendicular to the alignment of the nanoparticles. And the distance between the simulated Au@Ag NPs in FDTD was set at 1.72 nm. Results are in Fig. 7.

 figure: Fig. 7.

Fig. 7. (A) Statistical results of interparticle distance of Au@Ag NPs. (B) –(H) The simulation of electric field distributions in different possible aggregation states of Au@Ag NPs with FDTD solutions.

Download Full Size | PDF

Appendix D: Calculation method of the LOD [27]

The LOD is estimated without Pb2+. And the colorimetric and SERS signal is at least three times higher than the background (deionized water). The calibration curve for Pb2+ is as follows:

$$Y = A + K \times Log_{10}X$$
Among them, the meaning of Y in the SERS method is the SERS signal of 2-NT after adding Pb2+ into the probe; “X” is the final concentration of Pb2+. “A” and “K” are the variables obtained by linear regression of the least square root of the signal concentration curve, “A” is the intercept of the calibration curve, and “K” is the slope of the calibration curve.

The calculation method for the LOD of the calibration curve is:

$$LOD = {10^{[[(Y_b + 3SD)/Yb - A]/K]}}$$
“Yb” is the average SERS signal of the blank samples (deionized water is only added to the probe without adding any heavy metal ions), and “SD” is the standard deviation of the blank samples.

Appendix E: Calculation method of the SD

The SD of the blank signal was calculated according to the well-known formula:

$$SD = \sqrt {\left( {\frac{\textrm{1}}{{\textrm{n - 1}}} \times {{\sum\limits_{i = 1}^n {(Xi - Xaverage)} }^2}} \right)}$$
where, “n” is the total number of the blank sample. “Xi” is the “i” sample of the series of measurements. “Xaverage” is the average value of the SERS signals obtained for the specific series of identical samples repeated “n” times.

Funding

National Natural Science Foundation of China (Nos. 61805143); Shanghai Pujiang Program (Nos. 18PJ1408700); Shanghai Sailing Program (Nos. 19YF1435400).

Disclosures

The authors declare that there are no conflicts of interest related to this article.

References

1. A. R. Flegal and D. R. Smith, “Current Needs for Increased Accuracy and Precision in Measurements of Low Levels of Lead in Blood,” Environ. Res. 58(1–2), 125–133 (1992). [CrossRef]  

2. P. A. Meyer, T. Pivetz, T. A. Dignam, D. M. Homa, J. Schoonover, and D. Brody, “Surveillance for Elevated Blood Lead Levels Among Children — United States, 1997—2001,” MMWR Surveill Summ Morbidity and mortality weekly report. Surveillance summaries 52, 1–21 (2003).

3. A. K. De, Environmental Chemistry, 3rd ed. (Royal Society of Chemistry, 2016)

4. P. Chooto, P. Wararatananurak, and C. Innuphat, “Determination of trace levels of Pb(II) in tap water by anodic stripping voltammetry with boron-doped diamond electrode,” ScienceAsia 36(2), 150–156 (2010). [CrossRef]  

5. B. P. Lanphear, K. Dietrich, P. Auinger, and C. Cox, “Cognitive Deficits Associated with Blood Lead Concentrations <10 pg/dL in US Children and Adolescents,” Public Health Reports 115(6), 521–529 (2000). [CrossRef]  

6. W. H. Organization, Guidelines for Drinking Water Quality, vol. 1, 2nd ed.

7. F. Kummrow, F. F. Silva, R. Kuno, A. L. Souza, and P. V. Oliveira, “Biomonitoring method for the simultaneous determination of cadmium and lead in whole blood by electrothermal atomic absorption spectrometry for assessment of environmental exposure,” Talanta 75(1), 246–252 (2008). [CrossRef]  

8. D. Citak and M. Tuzen, “A novel preconcentration procedure using cloud point extraction for determination of lead, cobalt and copper in water and food samples using flame atomic absorption spectrometry,” Food Chem. Toxicol. 48(5), 1399–1404 (2010). [CrossRef]  

9. Ş Tokalıoğlu, T. Oymak, and Ş Kartal, “Coprecipitation of lead and cadmium using copper(II) mercaptobenzothiazole prior to flame atomic absorption spectrometric determination,” Microchim. Acta 159(1–2), 133–139 (2007). [CrossRef]  

10. V. L. Dressler, D. Pozebon, and A. J. Curtius, “Determination of heavy metals by inductively coupled plasma mass spectrometry after on-line separation and preconcentration,” Spectrochim. Acta, Part B 53(11), 1527–1539 (1998). [CrossRef]  

11. L. Zhang, Z. Li, X. Du, R. Li, and X. Chang, “Simultaneous separation and preconcentration of Cr(III), Cu(II), Cd(II) and Pb(II) from environmental samples prior to inductively coupled plasma optical emission spectrometric determination,” Spectrochim. Acta, Part A 86, 443–448 (2012). [CrossRef]  

12. H. R. Badiei, C. Liu, and V. Karanassios, “Taking part of the lab to the sample: On-site electrodeposition of Pb followed by measurement in a lab using electrothermal, near-torch vaporization sample introduction and inductively coupled plasma-atomic emission spectrometry,” Microchem. J. 108, 131–136 (2013). [CrossRef]  

13. L. Liang, F. Lan, S. Ge, J. Yu, N. Ren, and M. Yan, “Metal-Enhanced Ratiometric Fluorescence/Naked Eye Bimodal Biosensor for Lead Ions Analysis with Bifunctional Nanocomposite Probes,” Anal. Chem. 89(6), 3597–3605 (2017). [CrossRef]  

14. Y. Wen, C. Peng, D. Li, L. Zhuo, S. He, L. Wang, Q. Huang, Q. H. Xu, and C. Fan, “Metal ion-modulated graphene-DNAzyme interactions: design of a nanoprobe for fluorescent detection of lead(II) ions with high sensitivity, selectivity and tunable dynamic range,” Chem. Commun. 47(22), 6278–6280 (2011). [CrossRef]  

15. X. H. Zhao, R. M. Kong, X. B. Zhang, H. M. Meng, W. N. Liu, W. Tan, G. L. Shen, and R. Q. Yu, “Graphene-DNAzyme based biosensor for amplified fluorescence “turn-on” detection of Pb2+ with a high selectivity,” Anal. Chem. 83(13), 5062–5066 (2011). [CrossRef]  

16. L. Wang, H. X. Cao, Y. S. He, C. G. Pan, T. K. Sun, X. Y. Zhang, C. Y. Wang, and G. X. Liang, “Facile preparation of amino-carbon dots/gold nanoclusters FRET ratiometric fluorescent probe for sensing of Pb2+/Cu2+,” Sens. Actuators, B 282, 78–84 (2019). [CrossRef]  

17. B. Zhang and C. Wei, “Highly sensitive and selective detection of Pb(2+) using a turn-on fluorescent aptamer DNA silver nanoclusters sensor,” Talanta 182, 125–130 (2018). [CrossRef]  

18. X. X. Song, H. Fu, P. Wang, H. Y. Li, Y. Q. Zhang, and C. C. Wang, “The selectively fluorescent sensing detection and adsorptive removal of Pb(2+) with a stable [delta-Mo8O26]-based hybrid,” J. Colloid Interface Sci. 532, 598–604 (2018). [CrossRef]  

19. A. G. Memon, X. Zhou, Y. Xing, R. Wang, L. Liu, M. Khan, and M. He, “Label-free colorimetric nanosensor with improved sensitivity for Pb2 + in water by using a truncated 8–17 DNAzyme,” Front. Environ. Sci. Eng. 13(1), 12 (2019). [CrossRef]  

20. M. Rong, J. Li, J. Hu, A. Chen, W. Wu, and J. Lyu, “A highly sensitive and colorimetric biosensor based on magnetic nano-DNAzyme for detection of lead (II) ion in real water samples,” J. Chem. Technol. Biotechnol. 93(11), 3254–3263 (2018). [CrossRef]  

21. F. Sang, X. Li, Z. Zhang, J. Liu, and G. Chen, “Recyclable colorimetric sensor of Cr(3+) and Pb(2+) ions simultaneously using a zwitterionic amino acid modified gold nanoparticles,” Spectrochim. Acta, Part A 193, 109–116 (2018). [CrossRef]  

22. Y. Guo, Z. Wang, W. Qu, H. Shao, and X. Jiang, “Colorimetric detection of mercury, lead and copper ions simultaneously using protein-functionalized gold nanoparticles,” Biosens. Bioelectron. 26(10), 4064–4069 (2011). [CrossRef]  

23. O. Guselnikova, P. Postnikov, M. Erzina, Y. Kalachyova, V. Svorcík, and O. Lyutakov, “Pretreatment-free selective and reproducible SERS-based detection of heavy metal ions on DTPA functionalized plasmonic platform,” Sens. Actuators, B 253, 830–838 (2017). [CrossRef]  

24. O. Guselnikova, P. Postnikov, A. Trelin, V. Švorčík, and O. Lyutakov, “Dual Mode Chip Enantioselective Express Discrimination of Chiral Amines via Wettability-Based Mobile Application and Portable Surface-Enhanced Raman Spectroscopy Measurements,” ACS Sens. 4(4), 1032–1039 (2019). [CrossRef]  

25. M. S. Frost, M. J. Dempsey, and D. E. Whitehead, “Highly sensitive SERS detection of Pb 2+ ions in aqueous media using citrate functionalised gold nanoparticles,” Sens. Actuators, B 221, 1003–1008 (2015). [CrossRef]  

26. C. Fu, W. Xu, H. Wang, H. Ding, L. Liang, M. Cong, and S. Xu, “DNAzyme-based plasmonic nanomachine for ultrasensitive selective surface-enhanced Raman scattering detection of lead ions via a particle-on-a-film hot spot construction,” Anal. Chem. 86(23), 11494–11497 (2014). [CrossRef]  

27. Y. Shi, H. Wang, X. Jiang, B. Sun, B. Song, Y. Su, and Y. He, “Ultrasensitive, Specific, Recyclable, and Reproducible Detection of Lead Ions in Real Systems through a Polyadenine-Assisted, Surface-Enhanced Raman Scattering Silicon Chip,” Anal. Chem. 88(7), 3723–3729 (2016). [CrossRef]  

28. Q. Zou, X. Li, T. Xue, J. Zheng, and Q. Su, “SERS detection of mercury (II)/lead (II): A new class of DNA logic gates,” Talanta 195, 497–505 (2019). [CrossRef]  

29. A. N. Severyukhina, B. V. Parakhonskiy, E. S. Prikhozhdenko, D. A. Gorin, G. B. Sukhorukov, H. Mohwald, and A. M. Yashchenok, “Nanoplasmonic chitosan nanofibers as effective SERS substrate for detection of small molecules,” ACS Appl. Mater. Interfaces 7(28), 15466–15473 (2015). [CrossRef]  

30. A. Pallagi, E. G. Bajnoczi, S. E. Canton, T. Bolin, G. Peintler, B. Kutus, Z. Kele, I. Palinko, and P. Sipos, “Multinuclear complex formation between Ca(II) and gluconate ions in hyperalkaline solutions,” Environ. Sci. Technol. 48(12), 6604–6611 (2014). [CrossRef]  

31. Z. C. Zhang, G. Helms, S. B. Clark, G. Tian, P. L. Zanonato, and L. Rao, “Complexation of Uranium(VI) by Gluconate in Acidic Solutions: a Thermodynamic Study with Structural Analysis,” Inorg. Chem. 48(8), 3814–3824 (2009). [CrossRef]  

32. D. Sawyer, “Metal-gluconate complexes,” Chem. Rev. 64(6), 633–643 (1964). [CrossRef]  

33. G. Frens, “Controlled nucleation for the regulation of the particle size in monodisperse gold suspensions,” Nature (London), Phys. Sci. 241(105), 20–22 (1973). [CrossRef]  

34. R. Choudhury and T. K. Misra, “Gluconate stabilized silver nanoparticles as a colorimetric sensor for Pb 2+,” Colloids Surf., A 545, 179–187 (2018). [CrossRef]  

35. D.-K. Lim, I.-J. Kim, and J.-M. Nam, “DNA-embedded Au/Ag core–shell nanoparticles,” Chem. Commun. 129(42), 5312–5314 (2008). [CrossRef]  

36. R. A. Alvarez-Puebla, D. S. Dos Santos Jr., and R. F. Aroca, “Surface-enhanced Raman scattering for ultrasensitive chemical analysis of 1 and 2-naphthalenethiols,” Anal. Chem. 129(12), 1251–1256 (2004). [CrossRef]  

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. (A) The synthesis principle of Gluc/2-NT@Au@Ag NPs. (B) Schematic illustration of SERS and colorimetric determination for Pb2+.
Fig. 2.
Fig. 2. TEM images of (A) Au NPs. (B) Gluc@Au@Ag NPs. (C) Gluc/2-NT@Au@Ag NPs. (D) The aggregations of Gluc/2-NT@Au@Ag NPs in the presence Pb2+ of 7.5E-6 M. (E) The extinction spectra of the Gluc/2-NT@Au@Ag NPs suspension (without the addition of Pb2+) for 5 days. (F) The hydrodynamic size of the Gluc/2-NT@Au@Ag NPs suspension (without the addition of Pb2+) for 5 days. (G)The histogram of SERS intensity of Gluc/2-NT-@Au@Ag NPs at 1621 cm−1 band. (inset: SERS mapping image integrated at 1621 cm−1 band and the corresponding bright-field image). (H) A serials of SERS spectra for all mapping blocks in (G) (RSD of the intensities of 1621 cm−1 band is calculated to be 9.365%).
Fig. 3.
Fig. 3. (A) Extinction spectra of (a) Au NPs, (b) Gluc@Au@Ag NPs, (c) Gluc/2-NT@Au@Ag NPs and (d) Gluc/2-NT@Au@Ag NPs aggregations in the presence Pb2+ of 7.5E-6 M. (inset: Corresponding photographs (Sequencing from left to right: Au NPs, Gluc@Au@Ag NPs, Gluc/2-NT@Au@Ag NPs, the aggregations of Gluc/2-NT@Au@Ag NPs in the presence Pb2+)). (B) The SERS spectra of (a) Gluc/2-NT@Au@Ag NPs aggregations in the presence Pb2+ of 7.5E-6 M. (b) Gluc/2-NT@Au@Ag NPs and (c) Gluc@Au@Ag NPs.
Fig. 4.
Fig. 4. (A) Extinction spectra of Gluc/2-NT@Au@Ag NPs in presence of 1E-5 M Pb2+ over time. (B) Plot of absorbance of SPR band at 417 nm with time in presence of different concentration of Pb2+ ions (1E-6 M, 5E-6 M and 1E-5 M). (C) Extinction spectral change of Gluc@Au@Ag NPs in presence of Pb2+ (10−4–10−6 M). (D) Linear plot of absorbance of SPR band at 417 nm at different concentrations of Pb2+. The error bars indicate the standard deviation of 15 measurements.
Fig. 5.
Fig. 5. (A) Raman spectra of Gluc/2-NT@Au@Ag NPs in presence of Pb2+ (10−11–10−6 M). (B) Linear plot of Raman intensity of the 1621 cm−1 peak at various concentrations of Pb2+. The error bars indicate the standard deviation of 15 measurements.
Fig. 6.
Fig. 6. Selectivity test with various metal ions. (A) Histogram of absorbance of SPR band at 417 nm in the presence of different metal ions. (inset: Photographs of Gluc/2-NT@Au@Ag NPs in the presence of different metal ions of 10−5 M). (B) Extinction spectra of Gluc/2-NT@Au@Ag NPs in the presence of different metal ions. (C) Histogram of Raman intensity of the 1621 cm−1 band in the presence of different metal ions. The error bars indicate the standard deviation of 15 measurements. (D) Raman intensity of the 1621 cm−1 band in the presence of different metal ions of 10−6 M.
Fig. 7.
Fig. 7. (A) Statistical results of interparticle distance of Au@Ag NPs. (B) –(H) The simulation of electric field distributions in different possible aggregation states of Au@Ag NPs with FDTD solutions.

Tables (4)

Tables Icon

Table 1. Linear plot of absorption band at 417 nm and SERS intensity of the 1621 cm−1 band upon the addition of Pb2+ in different concentrations.

Tables Icon

Table 2. The comparison of the detection of Pb2+ with previously reported methods.

Tables Icon

Table 3. Determination of Pb2+ spiked in the real water sample

Tables Icon

Table 4. Observed fundamental vibrational wavenumbers of 2-naphthalenethiol

Equations (6)

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

Y = 0.31786 7513.53449 X
Y = 8997.46 9 + 706.525 × L og 10 X
Recovery ( % ) = [ ( C d / C s ) ] × 100 %
Y = A + K × L o g 10 X
L O D = 10 [ [ ( Y b + 3 S D ) / Y b A ] / K ]
S D = ( 1 n - 1 × i = 1 n ( X i X a v e r a g e ) 2 )
Select as filters


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