Au/Ag2S dimeric nanostructures for highly specific plasmonic sensing of mercury(Ⅱ)

Xinyi Liang Xu Du Ao Liu Zhixiong Cai Jingwen Li Maosheng Zhang Qingxiang Wang Jingbin Zeng

Citation:  Xinyi Liang, Xu Du, Ao Liu, Zhixiong Cai, Jingwen Li, Maosheng Zhang, Qingxiang Wang, Jingbin Zeng. Au/Ag2S dimeric nanostructures for highly specific plasmonic sensing of mercury(Ⅱ)[J]. Chinese Chemical Letters, 2023, 34(3): 107491. doi: 10.1016/j.cclet.2022.05.005 shu

Au/Ag2S dimeric nanostructures for highly specific plasmonic sensing of mercury(Ⅱ)

English

  • Environmental pollution and diseases caused by heavy metals and their compounds have been one of the major global public health concerns, especially mercury pollution [1-3]. Mercury is a highly toxic pollutant, which can get into the body via skin contact or diet and destroy the human central nervous system, immune system, kidney function, etc. In the meantime, it can be enriched and amplified in organisms through food chain [4, 5]. Hence the detection of the mercury content in environment is necessary for protecting human health.

    Conventional approaches for the detection of Hg2+ include atomic absorption spectrometry [6], atomic fluorescence spectroscopy [7-9], inductively coupled plasma mass spectrometry (ICP-MS) [10], and inductively coupled plasma atomic emission spectrometry (ICP-AES) [11], etc. These methods are always sensitive and selective, and can satisfy the qualitative or quantitative detection of mercury. However, most of them require costly and cumbersome instruments, skilled technicians, and time-consuming pretreatment processes, which greatly limits their application in real-time and on-site detection.

    Colorimetry has found broad application for on-the-spot detection of a wide range of analytes, because of its simple operation, visual readout and low instrument requirements [12]. In addition to organic dyes applied in conventional colorimetric approaches, plasmonic nanoparticles (NPs) have been widely designed as nanoprobes for colorimetric sensing, as a result of their tunable localized surface plasmon resonance (LSPR) effects [13]. The intensity and frequency of LSPR are strongly associated with the size, composition, morphology, the dielectric properties of the nanoparticles, and dielectric environment, which offers an opportunity to design specific nanoprobes for target analytes [14]. A common strategy is utilizing the analyte-induced assembly and dispersion of nanoparticles, which always require ligand modification that can form hydrogen bonds, metal-ligand complexes, etc. For example, a series of ligands containing thiols [15-17], DNA [18, 19], peptides [20, 21], proteins [22] and others [23, 24] have been modified onto the surface of plasmonic nanoparticles to develop Hg2+ nanosensors. These methods often require additional ligand synthesis, and most of them have insufficient selectivity making it difficult for the detection of Hg2+ in natural samples with complex matrices.

    Janus nanoparticles are a special class of particles with asymmetric structures and compositions. The two sides of nanoparticles have different properties, resulting in interface electron transfer and synergistic effects. They have found extensive applications in biomedicine [25], catalysis [26], biosensing [27], and other fields. Among numerous Janus nanoparticles, bimetallic nanoparticles with noble metal (Au and Ag) or metallide at its dimeric structure are considered to be ideal candidates as colorimetric nanoprobes, because their two active centers are both exposed, and their LSPR effect is more sensitive to the changes in morphology, size, composition and other properties [28]. Thus, engineering bimetallic nanoparticles with desirable dimeric structures is supposed to be essential for improving the sensitivity and selectivity of colorimetric method [28, 29].

    In this work, we attempted to develop a high-performance Hg2+ nanosensor by constructing an Au/Ag2S dimeric structure, which is composed of an Au core with tunable LSPR effect, and an Ag2S portion having exclusive response towards Hg2+. Ag2S is a stable and insoluble sulfide precipitate with minimum solubility product (Ksp (Ag2S) = 6.3 × 10−50), making it highly inert towards all anions. Theoretically, only Hg2+ can react with Ag2S to form HgS (Ksp (HgS) = 4 × 10−53). When Au/Ag2S NPs are exposed to Hg2+, Ag2S will be transformed to HgS, leading to the changes in the composition of the shells, thereby changing the LSPR effect. So the color and absorption changes which are associated with the Hg2+ concentration can be observed by naked eyes or monitored by UV–vis spectroscopy (Fig. 1). Thus, a naked-eye readout and spectral quantitation for Hg2+ concentration can be obtained. The specificity of the method was validated by successfully analyzing raw sewage samples without any pretreatment.

    Figure 1

    Figure 1.  Scheme of the method for the detection of Hg2+ using Au/Ag2S dimeric NPs.

    Tetrachloroauric acid, trisodium citrate, silver nitrate, iodine, and all the other inorganic compounds were bought from Sinopharm Chemical Reagent Co., Ltd. The silver ammonia solution (0.03 mol/L) was obtained by blending AgNO3 (0.1 mol/L, 6 mL), NH3·H2O (25%–28%, 1240 µL), and NaOH (3 mol/L, 650 µL). Stock solution of Hg(NO3)2 (1000 mg/L) was provided by TanMo Quality Testing Technology Co., Ltd. (Beijing, China). Sewage and tap water samples were collected from a sewage disposal exit in our campus and the lab, respectively. Deionized water was used for all experiments. UV–vis spectra were acquired from a spectrophotometer (UV-2450, Shimadzu). Transmission electron microscopy (TEM) images were captured by a JEM 1400 microscope (JEOL). High-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM), high-resolution transmission electron microscopy (HRTEM), and energy dispersive X-ray (EDX) elemental mapping measurements were conducted on a Tecnai F30 microscope. Powder X-ray diffraction (XRD) was measured by a X'Pert diffractometer (PANalytical).

    Citrate-stabilized Au NPs with the diameter of 13 nm were synthesized using a standard citrate method. Briefly, 10 mL of trisodium citrate (38.8 mmol/L) was injected into the boiling water (100 mL) containing 5 mL of HAuCl4 (24.28 mmol/L). After heating for 15 min, the faint yellow solution turned wine-red, which suggested Au NPs (13-nm diameter) had been successfully synthesized. Then, using the pre-made Au NPs as seeds, Au@Ag core-shell NPs were prepared via the epitaxial deposition of Ag on Au surface. In brief, 13 nm Au NPs (2 mL) were added into the mixture containing silver ammonia solution (0.3 mL), formaldehyde (HCHO) (0.48 mL, 0.01 mol/L) and ultrapure water (7.22 mL) and stirred for 10 min at 25 ℃. During the reaction, the colloid color gradually turned from wine-red to orange, suggesting the generation of Au@Ag core-shell NPs. To synthesize Au@Ag core-shell NPs which have different shell thicknesses, we can alter the ratio between HCHO and silver ammonia solution. Hence, we chose to fix the amount of HCHO, and adjusted the concentration of silver ammonia solution to alter the ratio. Afterwards, 0.5 mL of I2 was titrated into Au@Ag core-shell NPs (4 mL) and the mixture was stirred for 5 min, during which the colloid color varied from orange to purple-red. By changing the amount of I2, Au/AgI NPs with different AgI thickness were prepared to explore the influence on detection effect. Finally, Au/Ag2S NPs were acquired via the mix of Na2S (1.5 mL) and Au/AgI NPs (4.5 mL), whose color changed from purple-red to gray purple. After stirring for 10 min, a certain amount of PVP was added to reduce the risk of agglomeration. The supernatant was removed by centrifuging at 12,000 rpm for 10 min, redispersed in water. By varying Na2S concentration, the thickness effect of Ag2S on method sensitivity was studied.

    The Hg2+ standard solutions of different concentrations in the range of 0–154 µmol/L were mixed with the prepared Au/Ag2S NPs in a ratio of 1:9, and the UV–vis spectrophotometer was used for spectral determination. For assessing the practicability of Au/Ag2S NPs for natural water sample analysis, recovery experiments were carried out for detecting tap water and sewage samples. Firstly, we determined the Hg2+ concentrations in water samples. Then, known amounts of Hg2+ at different concentrations (30.81, 61.61 and 77.02 µmol/L) were spiked into these samples. Each sample was tested three times. According to the working curve equation, the Hg2+ content in different samples was obtained, and then the recovery rate and relative standard deviation (RSD) were calculated to evaluate the practicability of the method. The recovery was calculated by the formula R = (Ct-C0)/Cs × 100%, where Ct is the analytical result of the spiked sample, C0 is the result before spiking of the sample, and Cs is the amount of constituent added to the sample.

    Au NPs with a diameter of 13 nm were prepared by the standard citrate method, and then Ag shells were epitaxially grown onto the surface of Au NPs by silver mirror reaction. The uniform morphology could be observed on the basis of the analysis of the TEM image and the histogram of size distribution of Au@Ag core-shell NPs (Fig. S1 in Supporting information), and the average size is 21.3 ± 1.9 nm. After that, I2 was added to oxidize the Ag shell to generate Au/AgI dimeric NPs. It can be seen from Figs. 2a and b that Au/AgI NPs have a dimeric structure, which consists of Au centers with dark contrast and AgI shells with light contrast. HAADF-STEM and EDX elemental maps were utilized to characterize the chemical components and distribution of nanoparticles. Figs. 2c-f depict that Ag and I mostly overlap to form the AgI shell, and the shell mostly coated on the side of the Au core and only a thin layer on the other side, which further confirms the proposed dimeric structures. It is worth noting that AgI is susceptible to high voltages, making them easily decompose during the STEM test. Therefore, the mapping result shows partial aggregation. Such a dimeric nanostructure was formed due to the lattice-mismatch between Au and AgI as reported previously [28].

    Figure 2

    Figure 2.  (a) TEM image, (b) HRTEM image, (c, d) HAADF-STEM image, (e, f) EDX elemental maps of Au/AgI NPs. Orange line in (c) represents the EDX line scan shown in (f).

    To obtain the Au/Ag2S dimeric NPs, Na2S was used to etch and precipitate AgI shell into Ag2S. Typical TEM and HRTEM images in Figs. 3a and b illustrate that the resulting nanoparticles still maintain the dimeric structure, consisting of a dark Au core and an Ag2S attachment. The HAADF-STEM image and the EDX elemental maps (Figs. 3c and d) further confirm that the generated nanoparticles have the dimeric structure and the Au primarily located in the core, while Ag2S predominantly located at the side.

    Figure 3

    Figure 3.  (a) TEM image, (b) HRTEM image, (c) HAADF-STEM image, (d) EDX elemental maps of Au/Ag2S NPs.

    For revealing the colorimetric detection mechanism, the UV–vis spectra of the synthesized products in each step and the final detection products were characterized. As displayed in Fig. 4a, the LSPR band of Au NPs was centered at 520 nm and the solution was wine-red. Upon the addition of silver ammonia solution and HCHO, the absorption peak of sliver shell appeared at 392 nm, and the peak of Au blue-shifted from 520 nm to 494 nm due to the shielding effect by Ag [30]. Simultaneously, the colloid color turned to orange. Then, I2 was added in, and the solution gradually varied from orange to purple-red. A new band appeared at 421 nm, and it can be regarded as the band gap absorption of AgI [31]. Simultaneously, the absorption band of Au returned to 521 nm again, indicating that the shielding effect of sliver shell was weakened owing to the formation of Au/AgI NPs. This is in accordance with the characterization results in Fig. 2. Following the addition of Na2S to form Au/Ag2S dimeric NPs, the solution turned gray purple and the LSPR band at 421 nm vanished while the band at 521 nm moved to 562 nm. Such a redshift alludes to the transformation from AgI to Ag2S, giving rise to the enhancement in the refractive index of the media around the Au NPs [28, 32-34]. When a certain concentration of Hg2+ was mixed with the nanoprobes, a color change from gray purple to dark green and finally to navy occurred and the peak redshifted to 569 nm. The red-shift of spectrum can be attributed to the transformation of the composition from Ag2S to HgS with increasing refractive index from ~2.2 to ~2.9 [33-36].

    Figure 4

    Figure 4.  (a) UV–vis spectra of Au NPs (1), Au@Ag NPs (2), Au/AgI NPs (3), Au/Ag2S NPs (4), and Au/Ag2S NPs + Hg2+ (5). The inset is photograph of corresponding solutions. (b) TEM image, (c) HRTEM image, (d) HAADF-STEM image, (e, f) EDX elemental maps of Au/Ag2S NPs + Hg2+. Red line in d represents the EDX line scan shown in (f).

    To reveal the composition transformation, TEM and HAADF-STEM were utilized to characterize the nanoparticles generated from the reaction between Au/Ag2S dimeric NPs and Hg2+. Figs. 4b-d show that the generated products also possess a dimeric structure analogous to that of Au/Ag2S NPs. By means of lattice fringe analysis, we found (111) HgS with lattice spacing of 3.4 Å and (111) Au with lattice spacing of 2.4 Å in a single particle. To further figure out their elemental distribution, EDX elemental mapping measurements were employed. According to Fig. 4e, it can be seen that Au, as the core of the nanostructure, has a regular spherical shape, while S and Hg are basically in the same position, which can be inferred to form HgS and accumulate on the side of the core. Fig. 4f shows that Hg and S appeared simultaneously before Au in the line scan direction. Thus, the shell layer was transformed from Ag2S to HgS to form the Au/HgS dimeric NPs. In addition, Fig. 5 presents the XRD results of Au@Ag core-shell NPs, Au/AgI dimeric NPs, Au/Ag2S dimeric NPs, and Au/HgS dimeric NPs. These results validate the transformation of Ag, AgI, Ag2S and HgS in order.

    Figure 5

    Figure 5.  XRD results of (a) Au@Ag NPs, (b) Au/AgI NPs, (c) Au/Ag2S NPs, (d) Au/HgS NPs.

    The above results elucidate that the colorimetric detection mechanism depends on the fact that Hg2+ can react with Ag2S to constitute HgS, thereby varying the components and morphology of the nanomaterials and eliciting obvious spectral and color changes. The reaction between Au/Ag2S dimeric NPs and Hg2+ is shown below:

    (1)

    On account of the low solubility product of Ag2S, Au/Ag2S NPs are resistant to reacting with most interfering compounds, suggesting the great feasibility and high specificity of the colorimetric method for Hg2+ detection.

    On the basis of the above detection mechanism, the method sensitivity is highly relevant with the morphology and size of Au/Ag2S NPs. In order to obtain high detection efficiency, the core-to-shell proportion of Au@Ag NPs, the amount of I2 and the concentration of S2− were optimized.

    The concentration ratio between sliver ammonia solution and HCHO determines the core-to-shell proportion of Au@Ag NPs, which can further impact the shell thickness of Au/Ag2S NPs. Herein, we fixed the concentration of HCHO (0.01 mol/L), and varied the concentration of [Ag(NH3)2]+ from 0.06, 0.12, 0.24, 0.48, 0.72 to 0.96 mmol/L. When the concentration of [Ag(NH3)2]+ was low, the generated Ag shell was thin and the final Ag2S shell was also thin. After the addition of Hg2+, the spectral changes were minor and the sensitivity was low (Fig. S2 in Supporting information). For comparison, when the [Ag(NH3)2]+ with a higher concentration was added, a thicker Ag shell was generated, which had a greater shielding effect on the Au core. This led to a narrow range of spectral variations, which was also not conducive to visual observation and spectral analysis (Figs. S2c-f). Fig. S2b shows the resultant nanomaterials exhibited the highest sensitivity as the concentration of [Ag(NH3)2]+ is 0.12 mmol/L.

    Another crucial factor is the amount of I2 because it impacts on the conversion efficiency of Ag to AgI. Au/AgI NPs acquired by different amounts of I2 (0, 200, 500 and 700 µL) were used to form Au/Ag2S NPs for Hg2+ detection. Figs. S3a and b (Supporting information) demonstrates that insufficient I2 caused a narrow detection range of Hg2+ for the final synthetic Au/Ag2S NPs. This is probable due to the fact that insufficient I2 cannot completely oxidize Ag shell and the nanoparticles were still core-shell structure, which produced deficient Ag2S shell for Hg2+ sensing. Excessive I2 produced thick AgI shell, which had a strong shielding effect and reduced the synergistic effect, and compromised the sensitivity (Fig. S3d in Supporting information). Therefore, as displayed in Fig. S3c (Supporting information), we selected 500 µL I2 to obtain the nanoprobes.

    The influence of Ag2S attachments thickness on detection effect was also explored by transforming AgI to Ag2S with different S2− concentration (0, 0.2, 0.5, and 0.7 mmol/L). Based on the results provided in Fig. S4 (Supporting information), a small amount of S2− added caused inadequate conversion from AgI to Ag2S and affected the detection of Hg2+. While adding a large amount of S2−, the spectrum become irregular. Finally, we chose 0.2 mmol/L Na2S to synthesize Au/Ag2S NPs for Hg2+ detection.

    The performance of the developed method including limits of detection, reproducibility, linearity and selectivity was researched under the optimized conditions. As demonstrated in Fig. 6a, with increasing amount of Hg2+, an obvious color change from gray purple to dark green and finally to navy took place. Fig. 6b further shows that with Hg2+ concentration increasing from 0 to 100 µmol/L, the LSPR band red-shifted and the intensity showed a regular increase trend at 550 nm, which can be ascribed to the gradual conversion of Ag2S to HgS [35-37]. Fig. 6c shows that the variation of the peak intensity at 550 nm (∆A550) varied linearly with the concentration of Hg2+ and the equation is ∆A (550 nm) = 0.0041 × C[Hg2+] (µmol/L) + 0.0167 (R2 = 0.9894). The limit of detection (LOD) is 1.21 µmol/L, calculated by the signal-to-noise of 3. The reproducibility of this method was validated by testing six replicates of standard solutions with Hg2+ concentration of 30.8 µmol/L, and the RSD was 0.12%.

    Figure 6

    Figure 6.  (a) Photographs and (b) UV–vis spectra of Au/Ag2S NPs reacting with various Hg2+ concentrations. (c) The linear relationship between decreased absorbance at 550 nm and Hg2+ concentrations.

    To study the specificity of the proposed approach for Hg2+ detection, a number of anions and metal ions were detected under the parallel experimental conditions (Fig. 7). For the Hg2+ detection, the absorbance at 550 nm increased obviously with the color change from gray purple to dark green and finally to navy. While the addition of other compounds, even at the concentration of 25 to 1250 times of Hg2+, cannot produce apparent spectral and color change. It is worth noting that S2O32−, I and Br have intensive coordination capability and usually corrode Ag with the aid of oxygen, so they may interfere with the detection. Nevertheless, the results show that Au/Ag2S NPs exhibit a minimal response toward these ions, indicating this method has high selectivity and anti-interference ability. As mentioned before, such high selectivity is attributed to the extremely low solubility product of Ag2S and low reactivity with those interferents. Compared to other reported nanoprobes for Hg2+ detection shown in Table S1 (Supporting information), the proposed method shows comparable sensitivity and linear response range, but has better selectivity and eliminates the needs for surface modification, making it viable in the direct analysis of Hg2+ in complex samples. To evaluate the viability of Au/Ag2S NPs in natural water samples analysis, recovery tests were carried out for detecting tap water and sewage samples. As summarized in Table 1, the Hg2+ concentration in tap water was lower than the LOD of the approach, and the Hg2+ concentration in sewage samples was calculated as 15.40 µmol/L. The recoveries ranged from 100.5% to 102.8% and 93.1% to 97.3% in tap water and sewage samples, respectively, with RSDs in the range of 1.38%−2.89%, indicating the viability of the proposed method for Hg2+ in natural water samples. It should be emphasized that the sewage sample was directly analyzed without any sample pretreatment, indicating the high selectivity of the method.

    Figure 7

    Figure 7.  The ΔA550 of Au/Ag2S NPs reacting with Hg2+ or interfering compounds (K+, Na+, 30 mmol/L; Ca2+, Zn2+, Ba2+, Cd2+, Cu2+, Sn2+, Ni2+, Mn2+, Co2+, 2 mmol/L; Cr3+, Pb2+, Al3+, 1.5 mmol/L; Fe3+, 1 mmol/L) (a) and (NaCl, Cl, HCO3, CO32−, 50 mmol/L; HCl, HNO3, H2O2, IO32−, Br, I, OH, 25 mmol/L; ClO, F, PO43−, ClO4, S2O32−, 12.5 mmol/L; SO32−, BrO3, 5 mmol/L) (b).

    Table 1

    Table 1.  Analysis of Hg2+ in natural water samples.
    DownLoad: CSV

    To summarize, we proposed a rapid and highly specific colorimetric method for Hg2+ detection based on Au/Ag2S dimeric NPs. We synthesized Au/Ag2S dimeric nanoparticles via the epitaxial and lattice-mismatch process, which had an Au core with excellent optical properties and an Ag2S portion with exclusive response towards Hg2+. The sensing mechanism is attributed to the precipitation conversion of Ag2S with Hg2+ to generate more stable HgS. This alters the components and structure of the nanoparticles, further causing the change of the LSPR effect, and the color and spectrum will also change accordingly. Due to the minimum solubility product of Ag2S, Au/Ag2S dimeric NPs have a minimal response to common interferents, so this method has high selectivity towards Hg2+. Finally, the practicability of the method was validated by analyzing the raw sewage sample with satisfactory recoveries. This work not only provides a highly specific colorimetric method for on-site detection of Hg2+, but also offers an insight to the design of specific plasmonic nanostructures as unique optical nanoprobes.

    The authors declare no competing financial interest.

    This work was supported by the National Natural Science Foundation of China (No. 21876206), the Key Fundamental Project of Shandong Natural Science Foundation (No. ZR2020ZD13), the Science and Technology Projects of Qingdao (No. 21–1–4-sf-7-nsh), and the Youth Innovation and Technology project of Universities in Shandong Province (No. 2020KJC007).

    Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.cclet.2022.05.005.


    1. [1]

      S. Liu, X. Wang, G. Guo, Z. Yan, J. Environ. Manag.277 (2021) 111442. doi: 10.1016/j.jenvman.2020.111442

    2. [2]

      D.G. Streets, H.M. Horowitz, D.J. Jacob, et al., Environ. Sci. Technol. 51 (2017) 5969–5977. doi: 10.1021/acs.est.7b00451

    3. [3]

      C.T. Driscoll, R.P. Mason, H.M. Chan, D.J. Jacob, N. Pirrone, Environ. Sci. Technol. 47 (2013) 4967–4983. doi: 10.1021/es305071v

    4. [4]

      T.W. Clarkson, L. Magos, G.J. Myers, New Engl. J. Med. 349 (2003) 1731–1737. doi: 10.1056/NEJMra022471

    5. [5]

      K.H. Kim, E. Kabir, S.A. Jahan, J. Hazard. Mater. 306 (2016) 376–385. doi: 10.1016/j.jhazmat.2015.11.031

    6. [6]

      H. Erxleben, J. Ruzicka, Anal. l Chem. 77 (2005) 5124–5128. doi: 10.1021/ac058007s

    7. [7]

      Q. Niu, X. Wu, T. Li, et al., J. Fluoresc. 26 (2016) 1053–1058. doi: 10.1007/s10895-016-1793-4

    8. [8]

      J.V. Ros-Lis, M.D. Marcos, R. Mártinez-Máñez, K. Rurack, J. Soto, Angew. Chem. Inter. Ed. 44 (2005) 4405–4407. doi: 10.1002/anie.200500583

    9. [9]

      Y. Tang, F. He, M. Yu, F. Feng, L. An, et al., Macromolecul. Rapid Commun. 27 (2006) 389–392. doi: 10.1002/marc.200500837

    10. [10]

      J.A. Moreton, H. Trevor Delves, J. Anal. Atom. Spectr. 13 (1998) 659–665. doi: 10.1039/a802242i

    11. [11]

      Z. Zhu, G.C.Y. Chan, S.J. Ray, X. Zhang, G.M. Hieftje, Anal. Chem. 80 (2008) 7043–7050. doi: 10.1021/ac8011126

    12. [12]

      A. Piriya V. S, P. Joseph, K. Daniel, et al., Mater. Sci. Eng. C 78 (2017) 1231–1245. doi: 10.1016/j.msec.2017.05.018

    13. [13]

      N. Ullah, M. Mansha, I. Khan, A. Qurashi, TrAC Trends Anal. Chem. 100 (2018) 155–166. doi: 10.1016/j.trac.2018.01.002

    14. [14]

      B. Sepúlveda, P.C. Angelomé, L.M. Lechuga, L.M. Liz-Marzán, Nano Today 4 (2009) 244–251. doi: 10.1016/j.nantod.2009.04.001

    15. [15]

      D. Liu, W. Qu, W. Chen, et al., Anal. Chem. 82 (2010) 9606–9610. doi: 10.1021/ac1021503

    16. [16]

      C.J. Yu, W.L. Tseng, Langmuir 24 (2008) 12717–12722. doi: 10.1021/la802105b

    17. [17]

      F. Chai, C. Wang, T. Wang, Z. Ma, Z. Su, Nanotechnology 21 (2009) 025501.

    18. [18]

      C.W. Liu, Y.T. Hsieh, C.C. Huang, Z.H. Lin, H.T. Chang, Chem. Comm. (2008) 2242–2244. doi: 10.1039/b719856f

    19. [19]

      J.S. Lee, M.S. Han, C.A. Mirkin, Angew. Chem. Int. Ed. 46 (2007) 4093–4096. doi: 10.1002/anie.200700269

    20. [20]

      J.M. Slocik, J.S. Zabinski Jr, D.M. Phillips, R.R. Naik, Small 4 (2008) 548–551. doi: 10.1002/smll.200700920

    21. [21]

      J. Du, Y. Sun, L. Jiang, et al., Small 7 (2011) 1407–1411. doi: 10.1002/smll.201002270

    22. [22]

      Y. Guo, Z. Wang, W. Qu, H. Shao, X. Jiang, Biosens. Bioelectron. 26 (2011) 4064–4069. doi: 10.1016/j.bios.2011.03.033

    23. [23]

      J. Yang, Y. Zhang, L. Zhang, et al., Chem. Commun. 53 (2017) 7477–7480. doi: 10.1039/C7CC02198D

    24. [24]

      X. Chen, Y. Zu, H. Xie, A.M. Kemas, Z. Gao, Analyst 136 (2011) 1690–1696. doi: 10.1039/c0an00903b

    25. [25]

      G. Agrawal, R. Agrawal, A.C.S. Appl. Nano Mater. 2 (2019) 1738–1757. doi: 10.1021/acsanm.9b00283

    26. [26]

      T. Yang, L. Wei, L. Jing, et al., Angew. Chem. Int. Ed. 56 (2017) 8459–8463. doi: 10.1002/anie.201701640

    27. [27]

      P. Yánez-Sedeño, S. Campuzano, J.M. Pingarrón, Appl. Mater. Today 9 (2017) 276–288. doi: 10.1016/j.apmt.2017.08.004

    28. [28]

      J. Zeng, M. Li, A. Liu, et al., Adv. Funct. Mater. 28 (2018) 1800515. doi: 10.1002/adfm.201800515

    29. [29]

      A.S.D.S. Indrasekara, B.J. Paladini, D.J. Naczynski, et al., Adv. Healthcare Mater. 2 (2013) 1370–1376. doi: 10.1002/adhm.201200370

    30. [30]

      L. Rivas, S. Sanchez-Cortes, J.V. García-Ramos, G. Morcillo, Langmuir 16 (2000) 9722–9728. doi: 10.1021/la000557s

    31. [31]

      D.B. Pedersen, S. Wang, J. Phys. Chem. C 111 (2007) 1261–1267. doi: 10.1021/jp066357s

    32. [32]

      J. Yang, J.Y. Ying, Chem. Commun. (2009) 3187–3189. doi: 10.1039/b823320a

    33. [33]

      B. Xiong, R. Zhou, J. Hao, et al., Nat. Comm. 4 (2013) 1708. doi: 10.1038/ncomms2722

    34. [34]

      G. Park, C. Lee, D. Seo, H. Song, Langmuir 28 (2012) 9003–9009. doi: 10.1021/la300154x

    35. [35]

      F. Zhang, L. Zeng, C. Yang, et al., Analyst 136 (2011) 2825–2830. doi: 10.1039/c1an15113d

    36. [36]

      S. Chen, J. Tang, Y. Kuang, et al., Sensors Actuator. B: Chem 221 (2015) 1182–1187. doi: 10.1016/j.snb.2015.07.083

    37. [37]

      L.M. Liz-Marzán, Langmuir 22 (2006) 32–41. doi: 10.1021/la0513353

  • Figure 1  Scheme of the method for the detection of Hg2+ using Au/Ag2S dimeric NPs.

    Figure 2  (a) TEM image, (b) HRTEM image, (c, d) HAADF-STEM image, (e, f) EDX elemental maps of Au/AgI NPs. Orange line in (c) represents the EDX line scan shown in (f).

    Figure 3  (a) TEM image, (b) HRTEM image, (c) HAADF-STEM image, (d) EDX elemental maps of Au/Ag2S NPs.

    Figure 4  (a) UV–vis spectra of Au NPs (1), Au@Ag NPs (2), Au/AgI NPs (3), Au/Ag2S NPs (4), and Au/Ag2S NPs + Hg2+ (5). The inset is photograph of corresponding solutions. (b) TEM image, (c) HRTEM image, (d) HAADF-STEM image, (e, f) EDX elemental maps of Au/Ag2S NPs + Hg2+. Red line in d represents the EDX line scan shown in (f).

    Figure 5  XRD results of (a) Au@Ag NPs, (b) Au/AgI NPs, (c) Au/Ag2S NPs, (d) Au/HgS NPs.

    Figure 6  (a) Photographs and (b) UV–vis spectra of Au/Ag2S NPs reacting with various Hg2+ concentrations. (c) The linear relationship between decreased absorbance at 550 nm and Hg2+ concentrations.

    Figure 7  The ΔA550 of Au/Ag2S NPs reacting with Hg2+ or interfering compounds (K+, Na+, 30 mmol/L; Ca2+, Zn2+, Ba2+, Cd2+, Cu2+, Sn2+, Ni2+, Mn2+, Co2+, 2 mmol/L; Cr3+, Pb2+, Al3+, 1.5 mmol/L; Fe3+, 1 mmol/L) (a) and (NaCl, Cl, HCO3, CO32−, 50 mmol/L; HCl, HNO3, H2O2, IO32−, Br, I, OH, 25 mmol/L; ClO, F, PO43−, ClO4, S2O32−, 12.5 mmol/L; SO32−, BrO3, 5 mmol/L) (b).

    Table 1.  Analysis of Hg2+ in natural water samples.

    下载: 导出CSV
  • 加载中
计量
  • PDF下载量:  3
  • 文章访问数:  235
  • HTML全文浏览量:  20
文章相关
  • 发布日期:  2023-03-15
  • 收稿日期:  2022-01-15
  • 接受日期:  2022-05-05
  • 修回日期:  2022-04-16
  • 网络出版日期:  2022-05-10
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

/

返回文章