Green synthesis of MIL-101/Au composite particles and their sensitivity to Raman detection of thiram

Huihui LIU Baichuan ZHAO Chuanhui WANG Zhi WANG Congyun ZHANG

Citation:  Huihui LIU, Baichuan ZHAO, Chuanhui WANG, Zhi WANG, Congyun ZHANG. Green synthesis of MIL-101/Au composite particles and their sensitivity to Raman detection of thiram[J]. Chinese Journal of Inorganic Chemistry, 2024, 40(10): 2021-2030. doi: 10.11862/CJIC.20240059 shu

MIL-101/Au复合粒子的绿色合成及其对福美双的拉曼检测灵敏性

    通讯作者: 王智, shikouri@163.com
    张丛筠, zhangcy@qdu.edu.cn
  • 基金项目:

    山东省自然科学基金 ZR2022MB141

    国家自然科学基金 51701186

摘要: 采用静电自组装方式构建了金属有机骨架(MOF) MIL-101和具有表面等离激元(SPP)的金纳米粒子(Au NPs)的复合材料, 该材料可作为高灵敏度、可重复使用的表面增强拉曼散射(SERS)检测平台。在水溶液中, 无需任何修饰剂, 仅利用前体粒子的电负性, 便成功制备了稳定的复合粒子。由于MIL-101的富集能力和Au NPs的电磁增强效应, 设计的MIL-101/Au复合粒子具有超高的SERS灵敏度, 对罗丹明6G(R6G)的检出限低至10-10 mol·L-1。同时, 衬底具有出色的稳定性、良好的再现性和可回收性。此外, 该基底可用于直接捕获和灵敏检测福美双等农药残留。

English

  • In recent years, the growing problem of abuse of pesticides is a serious threat to the human health and ecological environment. Therefore, rapid and accurate analysis of pesticide residues in food detection has been developed[1-3]. The conventional analytic methods for pesticide residues mainly include gas chromatography (GC), high‑performance liquid chromatography (HPLC), chromatography-mass spectrometry, enzyme-linked immunoassay (ELISA), and enzyme inhibition[4-7]. However, these methods generally need sophisticated instrumentation, time-consuming, and complicated operations, which limit their application for rapid detection.

    Surface-enhanced Raman scattering (SERS) spectroscopy emerging as a non-destructive and sensitive (even at the molecular level) technique can offer the chemical structure and molecular composition fingerprint information[8-10]. During the past decades, SERS has gained increasing interest in material science because of its tremendous applications in environmental chemistry, medical detection, and food safety[11-13]. To date, there are mainly two types of enhancement mechanisms including electromagnetic enhancement (EM) and chemical enhancement (CM)[14-15], where the pronounced SERS effect is generated by local surface plasmon resonance (LSPR) and charge transfer. Generally, LSPR-induced SERS effects largely depend on the geometrical shape of nanomaterials and the gap size between plasmonic nanostructures[16]. At these "hot spots", several orders of magnitude could amplify the Raman scattering cross-section of adsorbed target molecules[17]. Therefore, it is desirable and challenging to construct abundant hot spots and improve the adsorption capability of molecules (especially for nonadsorbing organic pollutant molecules) of substrates.

    Among various materials, plasmonic materials with ultrahigh specific surface areas and enrichment ability could provide a great opportunity to rapidly absorb target molecules with weak affinity on metal surfaces and achieve highly effective SERS enhancement and detection[18-21]. Metal-organic frameworks (MOFs) are expected to emerge as attractive materials with many unique properties, such as high porosity, high specific surface areas, and tunable structure[22]. MOF-based SERS substrate can show high sensitivity, excellent selectivity, and outstanding stability[23-26]. Accordingly, MOF-based hybrids provide an opportunity to improve the signal enhancement ability.

    MOFs play crucial roles in MOF-derived metal SERS substrate. (ⅰ) A single-component MOF matrix could be used as an SERS base. Sun et al. demonstrated that the SERS enhancement factors (EFs) of MOF substrate with high tailorability could be greatly improved to 106 and the detection limit is as low as 10-8 mol·L-1 [22]. (ⅱ) The MOF shell can avoid direct contact between the plasmonic nanoparticles (PNPs) and the external environment, thus improving the stability of the substrate and extending the service life. Sun et al. constructed a highly sensitivity SERS sensor based on magnetic MOF. The Raman signal intensity of the substrate material was stable after 22 d of storage at room temperature[27]. (ⅲ) By employing the most convenient "seed growth method", Au@MIL-100, Au@MIL-101, Au@Cu3(BTC)2, PNPs@ZIF-8, and Au@AgNRs@ZIF-8 substrates have been successfully prepared for various SERS applications[28-32]. However, the easy in situ methods make it difficult to obtain uniform PNPs with reproductive signals[33].

    In this work, we have designed a SERS-active substrate by electrostatic self-assembly strategy without complicated surface modification. The composite particles can maintain a stable colloidal form in a water environment. The conjunction of MOF and the highly electromagnetic response of PNPs endow substrates with significantly enhanced Raman sensitivity, high reproducibility, and exceptional molecule harvesting capability. The resultant MIL-101/Au composites reveal sensitive SERS performance with a Rhodamine 6G (R6G) detection limit down to 10-10 mol·L-1. Only simple pretreatment is required, and the residue signal can be efficiently obtained even at trace concentrations. Of particular note, the MOF-based substrate could be a dual-functional SERS platform for the simultaneous capture and sensitive detection of pesticide residues.

    Chromium (Ⅲ) nitrate nonahydrate (Cr(NO3)3·9H2O), terephthalic acid, and tetramethylthiuram disulfide (TMTD) were purchased from Aladdin. Sodium borohydride (NaBH4), sodium citrate, and N, N-dimethylformamide (DMF) were obtained from Tianjin Damao Chemical Reagent. Hydrogen tetrachloroaurate(Ⅲ) trihydrate (HAuCl4·3H2O, 99.9%) was purchased from Macklin. Ultrapure water (Millipore, 18.2 MΩ·cm) was used in all experiments.

    MIL-101 NPs were synthesized by a typical method without modification. Specifically, 1 g Cr(NO3)3·9H2O and 0.9 g terephthalic acid were introduced into 15 mL deionized water. After 30 min of ultrasonication, the mixture was transferred to a 50 mL polytetrafluoroethylene lining. Then 90 μL HF was added into the mixture. After reaction for 8 h under 220 ℃, the product was collected by centrifugation, washed with ethanol, and dried in a vacuum under 150 ℃. The MIL-101 powder was dispersed in deionized water to obtain 1 mg·mL-1 MIL-101 dispersion liquid.

    The MIL-101/Au NPs were synthesized in a two-step process. The Au NPs were synthesized using a seed-growth method. Briefly, ice-cold NaBH4 solution (0.1 mol·L-1, 0.6 mL) was added into an aqueous mixture of HAuCl4 (5.0×10-4 mol·L-1, 10 mL) and sodium citrate (5.0×10-4 mol·L-1, 10 mL), followed by rapid stirring for 2 min. The seed solution was kept undisturbed for about 1 h at room temperature. Next, 600 μL Au seed dispersion was added into the growth solution containing HAuCl4 (5.0×10-4 mol·L-1, 50 mL) and sodium citrate (5.0×10-4 mol·L-1, 50 mL), respectively, to obtain Au NPs. The resulting solution was gently stirred for 30 min under 80 ℃ and then left undisturbed overnight. After centrifugation and washing with water twice, the Au NPs were re-dispersed in 20 mL water. MIL-101/Au NPs were prepared by mixing different volume ratios of MIL-101 dispersion liquid and Au NPs dispersion liquid solution (1∶1, 1∶2, 1∶3). After gentle vibration for 120 s, the mixture was undisturbed at room temperature for 1 h.

    The morphologies and structures of the prepared MIL-101/Au NPs were observed by field emission scanning electron microscopy (SEM, Hitachi SU-8010) at an accelerating voltage of 10 kV and by transmission electron microscopy (TEM, JEOL-2100F) at an accelerating voltage of 200 kV. The crystal structure of the synthesized NPs was analyzed by X‑ray diffraction (XRD, Haoyuan 2700B) equipped with Cu (λ=0.154 6 nm), and the diffraction patterns were collected on an X-ray diffractometer in the range of 5°-70° with a scanning rate of 2 (°)·min-1 at 40 V voltage and 40 mV current. Fourier transform infrared spectra (FTIR) were tested by a Nicolet iS50 spectrometer (Thermo Scientific). The ζ potential of different samples was investigated using an ζ potential analyzer (Zeta sizer NANO, ZS90). The Raman spectra were acquired from a confocal microscopy Raman spectrometer (Renishaw inVia) using a 532 nm laser as the excitation source.

    The prepared MIL-101/Au aqueous suspensions (0.2 mL) were injected into 0.8 mL R6G aqueous solutions with various concentrations (10-6, 10-7, 10-8, 10-9, 10-10 mol·L-1) and stirred for several minutes. Then 0.8 mL of supernatant was removed by centrifugation (4 000 r·min-1). The NPs were re-dispersed by ultrasound for 30 s. The mixed solutions were spin-coated at 500 r·min-1 on the silicon wafer and dried under ambient conditions for SRES measurement. The acquisition time was 10 s, and a 532 nm laser with 5 mW power was used as the excitation source.

    After the first recording of the SERS signal according to the above method, the substrate was detached from the wafer by sonication and then washed at reflux in 5 mL of ethanol solution for 20 min. After separating the MIL-101/Au particles from the ethanol, they were re-dispersed in 0.2 mL ultrapure water. The washed MIL-101/Au was again mixed with 10-6 mol·L-1 R6G solution (0.8 mL) for the next detection to obtain the SERS signal. This detection-washing procedure was repeated to test the recyclability of the MIL-101/Au substrate.

    The fabrication process of MIL-101/Au is illustrated in Fig. 1a. Firstly, MIL-101 NPs were synthesized via a hydrothermal method, while Au NPs were prepared using the seed-mediated growth method. The ζ potential on the surface of MIL-101 was measured using a nano-size analyzer, and the potential of MIL-101 in aqueous solution was about 29.7 mV, implying that MIL-101 is surrounded by a layer of positive charge. Meanwhile, due to the surface of Au NPs being modified by citrate anion, Au NPs are negatively charged. This is also evidenced by the value (-31.2 mV) from the ζ potential measurement. Then, the two particles were physically combined through electrostatic attraction, and Au NPs were anchored on the MIL-101 surface.

    Figure 1

    Figure 1.  Fabrication process of MIL-101/Au (a); SEM images of MIL-101 (b) and MIL-101/Au (c); TEM images of MIL-101 (d) and MIL-101/Au (e); FTIR spectra (f) and XRD patterns (g) of MIL-101 and MIL-101/Au

    As seen in Fig. 1b and 1d, the MIL-101 with the uniform size is approximately 300 nm, the particle size distribution is shown in Fig.S1 (Supporting information). Meanwhile, as shown in Fig. 1c and 1e, AuNPs are evenly distributed on the MIL-101 surface. The as-prepared substrate was further characterized using FTIR and XRD. As shown in Fig. 1f, for MIL-101 and MIL-101/Au samples, the two kinds of NPs show the same infrared absorption curve apart from the distinct difference in intensity. This indicates that the combination of MIL-101 and Au does not involve chemical interaction, and the weakening in the combined peak strength is attributed to the reduction of the relative content of MIL-101. The XRD patterns of MIL-101 and MIL-101/Au are shown in Fig. 1g. Diffraction peaks from 5° to 20° are attributed to MIL-101, while the weak peaks of MIL-101/Au at a small angle can be observed. Moreover, diffraction peaks of MIL-101/Au at 38.2°, 44.4°, and 64.6° are attributed to Au(111), (200), and (220), respectively. The formation of tiny peaks at small angles is the weakening effect caused by the high‑intensity Au peak. All the characterization methods above indicate the successful formation of Au-decorated MIL-101 NPs.

    It is well known that the plasma coupling effect of metal NPs depends largely on distribution density[34]. When two or more nanostructures approach each other (< 10 nm), the electric field between the nanostructures could be amplified greatly, generating a prominent "hot spots" effect that enhances the Raman signal. In this paper, the SERS performance of the composite substrate was optimized by changing the density of Au NPs loaded on the surface of MIL-101. Fig. 2a-2c shows the morphologies of MIL-101/Au NPs compounded with volume ratios of 1∶1, 1∶2, and 1∶3, respectively. When the added concentration of Au particles increases to a certain extent (1∶2, 1∶3), there is no significant difference in Au density on MIL-101.

    Figure 2

    Figure 2.  SEM images of the composite substrates of MIL-101/Au with volume ratios of 1∶1 (a), 1∶2 (b), and 1∶3 (c)

    To evaluate the SERS properties of MIL-101/Au, R6G was utilized as a Raman probe molecule. Fig. 3a shows the SERS spectra of 10-6 mol·L-1 R6G on the enhanced substrates prepared with three different covering densities of particles. Fig. 3b shows the Raman intensity of peaks at 1 506 cm-1 in Fig. 3a. After the volume ratio of MIL-101 dispersion liquid and HAuCl4 solution increased from 1∶1 to 1∶2, the Raman signal was significantly enhanced. However, when the amount of added Au continued to increase, the enhancement of the Raman signal was almost unchanged. Theoretically, increasing the density of Au NPs is beneficial for creating hot spots. However, for electrostatic interactions, potential balance and load saturation are limited. Thermogravimetric (TG) tests were performed to prove the amount of Au particles in the composite particles. It could be known from Fig. 3c that the carbon residue of pure MIL-101 at 700 ℃ is 31.4%. Au NPs remain stable at 700 ℃ and their mass does not change. Thus, the amount of carbon residue increased with the increase of Au particle loading. Moreover, when the load of Au particles reaches saturation turntable, the residual carbon value no longer increases. So the particles of MIL-101 and Au in volume ratio at 1∶2 were used as SERS substrates for subsequent performance tests.

    Figure 3

    Figure 3.  SERS spectra of 10-6 mol·L-1 R6G molecules at different volume ratios of MIL-101 dispersion liquid and HAuCl 4 solution as SERS substrates (a) and the corresponding Raman intensity at 1 506 cm-1 (b); TG curves of three composite substrates and pure MIL-101 (c)

    To explore the Raman enhancement mechanism of the substrate, 10-6 mol·L-1 R6G molecules on pure silicon, MIL-101, and MIL-101/Au composite particles were tested. Results are shown in Fig.S2, there was almost no Raman signal of R6G on the pure silicon wafer. For MIL-101, only characteristic peaks attributed to the base itself can be observed, although its high specific surface and enrichment properties, signal peaks of R6G molecule enrichment were hardly observed. Furthermore, the signal of R6G is significantly enhanced after MIL-101 is combined with Au, which combines the characteristics of MOF material and the LSPR of precious metal particles.

    The adsorption capability of MIL-101/Au composite substrates is also crucial for the enhancement of SERS signals. As shown in Fig. 4a and 4b. With the increase of adsorption time, the characteristic signal of R6G was gradually enhanced. After 40 min of adsorption, there was no more significant improvement in signal intensity, implying that the substrate reached adsorption saturation for R6G within 40 min.

    Figure 4

    Figure 4.  SERS spectra of 10-6 mol·L-1 R6G molecules with MIL-101/Au at different adsorption times (a) and the corresponding Raman intensity at 1 506 cm-1 (b)

    To explore the sensitivity of SERS sensing, R6G molecules with different concentrations were selected as target molecules to evaluate the limit of detection (LOD) of prepared substrates. As shown in Fig. 5a, when the concentration of R6G was as low as 10-10 mol·L-1, a weak SERS signal could still be observed. The combination of MIL-101 enrichment and LSPR enhancement of Au has a strong synergistic effect for trace detection. Meanwhile, the Raman peak strength and the logarithm of the concentration at the range of 10-5-10-10 mol·L-1 follow a good linear dependence. Such linear dependence enables SERS measurements to be used for quantitative analysis of low-concentration molecules.

    Figure 5

    Figure 5.  SERS spectra of R6G with different concentrations on the MIL-101/Au substrates (a) and calibration plot based on Raman intensity at 1 506 cm-1 (b)

    Apart from sensitivity, signal reproducibility, which is largely dependent on the uniformity of hotspot distribution, is another prerequisite for enhancing SERS performance and practical application. As shown in Fig. 6a, the Raman spectra of R6G with a concentration of 10-6 mol·L-1 from 20 points on the substrate were irregularly selected to evaluate the uniformity of SERS activity of MIL‑101/Au substrates. Fig. 6b reveals that the signal intensity variations are within 7% at the characteristic peak of 1 506 cm-1, indicating excellent uniform spatial distribution of hot spots and outstanding signal reproducibility.

    Figure 6

    Figure 6.  SERS spectra of 10-6 mol·L-1 R6G collected from 20 different points on the MIL-101/Au substrates (a) and the corresponding Raman intensity distribution at 1 506 cm-1 (b)

    The recyclability of SERS substrates is an effective way to reduce substrate costs and avoid secondary contamination. The recycling of MIL‑101/Au substrates was also investigated in this paper. It can be seen from Fig. 7 that MIL-101/Au substrates can be reused five times without significant attenuation of Raman strength, and there was no obvious signal after the substrate was washed during each cycle.

    Figure 7

    Figure 7.  SERS spectra of 10-6 mol·L-1 R6G with MIL-101/Au obtained during 1-5 detection-washing cycles

    To test the stability of synthetic MIL-101-based substrates, time-stable Raman measurements were performed by comparing the signal intensity of 10-6 mol·L-1 R6G on freshly prepared substrates and exposure to ambient conditions for 0, 5, 15, and 30 d. As time proceeded, the SERS signal intensity on the MIL-101/Au substrate continued to decay, with a weakening trend after 15 d, and retaining 79% of the initial signal intensity after 30 d (Fig.S3). This may be due to the collapse of part of the MOF structure in the environment, leading to a decrease in adsorption capacity. This phenomenon also indicates the relatively good stability of the substrate.

    TMTD is a kind of insecticide widely used in crops. The bioactive site of this kind of compound is the dithiocarbamate group, which can interact with enzymes containing sulfhydryl group (—SH) and coenzymes of fungus cells. It can penetrate human skin and mucous membranes, wherein the derivative has the possibility of causing cancer and human deformity. To investigate the MIL-101/Au-based SERS substrates for practical applications, various amounts of TMTD were spiked into acetone and the Raman spectra are displayed in Fig. 8. The lowest directly detectable concentration of thiram in acetone on the MIL-101/Au-based SERS substrates waw as low as 10-10 mol·L-1, further suggesting that this SERS substrate has a wide application prospect in the agricultural products detection field.

    Figure 8

    Figure 8.  SERS spectra of thiram with different concentrations on the MIL-101/Au substrates (a) and calibration plot based on Raman intensity at 1 378 cm-1 (b)

    In summary, a new SERS sensor composed of MOFs and Au NPs has been successfully fabricated via an eco-friendly assembly method. The Au NPs are loaded on the surface of MIL-101 at a suitable density via electrostatic effects to generate intense "hot spots" field. The nanomaterial composed of MOFs can concentrate trace target molecules into "hot spots" regions originating from Au NPs. Furthermore, the uniform distribution of Au NPs on the surface of MIL-101 can optimize the uniformity of the substrate. Ultimately, the fabricated MIL-101/Au composites were used as SERS sensors to detect trace TMTD in acetone. The results indicate that the LOD of this substrate can be as low as 10-10 mol·L-1 and the linear concentration range is from 10-5 to 10-10 mol·L-1 with a linearly dependent coefficient R2=0.993. This work provides a promising strategy to prepare SERS substrates and has great utilization potential in the fields of food and agricultural safety. Meanwhile, there is a complicated process of residual probe molecules handling during repeated detection for the substrate. Therefore, intervening particles with a photocatalytic effect can be considered to realize a convenient light reaction to remove the probe molecules.


    Acknowledgments: This work was supported by the Natural Science Foundation in Shandong Province of China (Grant No.ZR2022MB141) and the National Natural Science Foundation of China (Grant No.51701186). Supporting information is available at http://www.wjhxxb.cn Supporting information is available at http://www.wjhxxb.cn
    1. [1]

      Bernat A, Samiwala M, Albo J, Jiang X Y, Rao Q C. Challenges in SERS-based pesticide detection and plausible solutions[J]. J. Agric. Food. Chem., 2019, 67(45):  12341-12347. doi: 10.1021/acs.jafc.9b05077

    2. [2]

      Zhao P N, Liu H Y, Zhang L N, Zhu P H, Ge S G, Yu J H. Paper-based SERS sensing platform based on 3D silver dendrites and molec-ularly imprinted identifier sandwich hybrid for neonicotinoid quantifi-cation[J]. ACS Appl. Mater. Interfaces, 2020, 12(7):  8845-8854. doi: 10.1021/acsami.9b20341

    3. [3]

      Jin X, Zhu Q Y, Feng L, Li X, Zhu H Y, Miao H Y, Zeng Z F, Shi G. Light-trapping SERS substrate with regular bioinspired arrays for detecting trace dyes[J]. ACS Appl. Mater. Interfaces, 2021, 13(9):  11535-11542. doi: 10.1021/acsami.1c00702

    4. [4]

      Li G J, Zhang X, Liu T T, Fan H X, Liu H C, Li S Y, Wang D W, Ding L. Dynamic microwave-assisted extraction combined with liquid phase microextraction based on the solidification of a floating drop for the analysis of organochlorine pesticides in grains followed by GC[J]. Food Sci. Human Wellness, 2021, 10(3):  375-382. doi: 10.1016/j.fshw.2021.02.029

    5. [5]

      Birader K, Kumar P, Tammineni Y, Barla J A, Barla R, Suman P. Col-orimetric aptasensor for on-site detection of oxytetracycline antibiotic in milk[J]. Food Chem., 2021, 356:  129659. doi: 10.1016/j.foodchem.2021.129659

    6. [6]

      Park E, Lee J, Lee J, Lee J, Lee H S, Shin Y, Kim J H. Method for the simultaneous analysis of 300 pesticide residues in hair by LC-MS/MS and GC-MS/MS, and its application to biomonitoring of agricultural workers[J]. Chemosphere, 2021, 277:  130215. doi: 10.1016/j.chemosphere.2021.130215

    7. [7]

      Chen Z J, Wu H L, Xiao Z L, Fu H J, Shen Y D, Luo L, Wang H, Lei H T, Xu Z L. Rational hapten design to produce high-quality antibod-ies against carbamate pesticides and development of immunochro-matographic assays for simultaneous pesticide screening[J]. J. Hazard. Mater., 2021, 412:  125241. doi: 10.1016/j.jhazmat.2021.125241

    8. [8]

      Liebel M, Calderon I, Pazos-Perez N, Hulst N F V, Alvarez-Puebla R A. Widefield SERS for high-throughput nanoparticle screening[J]. Angew. Chem. Int. Ed., 2022, 61(20):  e202200072. doi: 10.1002/anie.202200072

    9. [9]

      Zhu W Q, Crozier K B. Quantum mechanical limit to plasmonic enhancement as observed by surface-enhanced Raman scattering[J]. Nat. Commun., 2014, 5:  5228. doi: 10.1038/ncomms6228

    10. [10]

      刘元君, 叶芬, 王威, 张俊豪, 晏超, 袁爱华. 蜂窝状Ag纳米颗粒膜制备及应用于表面增强拉曼散射基底[J]. 无机化学学报, 2019,35,(10): 1861-1868. LIU Y J, YE F, WANG W, ZHANG J H, YAN C, YUAN A H. Fabri-cation of honeycomb-like Ag nanoparticles film used as surface enhanced Raman scattering substrate[J]. Chinese J. Inorg. Chem., 2019, 35(10):  1861-1868.

    11. [11]

      Langer J, Jimenez de Aberasturi D, Aizpurua J. Present and future of surface-enhanced Raman scattering[J]. ACS Nano, 2019, 14(1):  28-117.

    12. [12]

      Luo X J, Zhao X J, Wallace G Q, Wallace M H, Wilkinson K J, Wu P, Cai C X, Bazuin C J, Masson J F. Multiplexed SERS detection of microcystins with aptamer-driven core-satellite assemblies[J]. ACS Appl. Mater. Interfaces, 2021, 13(5):  6545-6556. doi: 10.1021/acsami.0c21493

    13. [13]

      Zhang D J, You H J, Yuan L, Hao R, Li T, Fang J X. Hydrophobic slippery surface-based surface-enhanced Raman spectroscopy plat-form for ultrasensitive detection in food safety applications[J]. Anal. Chem., 2019, 91(7):  4687-4695. doi: 10.1021/acs.analchem.9b00085

    14. [14]

      Lin J J, Liang L B, Ling X, Zhang S Q, Mao N N, Zhang N, Sumpter B J, Meunier V, Tong L M, Zhang J. Enhanced Raman scattering on in-plane anisotropic layered materials[J]. J. Am. Chem. Soc., 2015, 137(49):  15511-15517. doi: 10.1021/jacs.5b10144

    15. [15]

      Ding S Y, You E M, Tian Z Q, Moskovits M. Electromagnetic theo-ries of surface-enhanced Raman spectroscopy[J]. Chem. Soc. Rev., 2017, 46(13):  4042-4076. doi: 10.1039/C7CS00238F

    16. [16]

      刘晓宇, 张东杰, 张慧娟, 张丛筠, 刘亚青. 金@银核壳纳米粒子的制备和形貌的精确控制及其表面增强拉曼光谱性能[J]. 无机化学学报, 2018,34,(4): 712-718. LIU X Y, ZHANG D J, ZHANG H J, ZHANG C Y, LIU Y Q. Synthesis of Au@Ag core-shell nanoparticles for sensitive surface-enhanced Raman scattering by precisely adjust its morphology[J]. Chinese J. Inorg. Chem., 2018, 34(4):  712-718.

    17. [17]

      Zhang K G, Yao S, Li G K, Hu Y L. One-step sonoelectrochemical fabrication of gold nanoparticle/carbon nanosheet hybrids for effi-cient surface-enhanced Raman scattering[J]. Nanoscale, 2015, 7(6):  2659-2666. doi: 10.1039/C4NR07082H

    18. [18]

      Koh C S L, Sim H Y F, Leong S X, Boong S K, Chong C, Ling X Y. Plasmonic nanoparticle-metal-organic framework (NP-MOF) nanohy-brid platforms for emerging plasmonic applications[J]. ACS Mater. Lett., 2021, 3(5):  557-573. doi: 10.1021/acsmaterialslett.1c00047

    19. [19]

      Lai H S, Li G K, Xu F G, Zhang Z M. Metal-organic frameworks: Opportunities and challenges for surface-enhanced Raman scatter-ing-A review[J]. J. Mater. Chem. C, 2020, 8(9):  2952-2963. doi: 10.1039/D0TC00040J

    20. [20]

      Zhang Y, Xue C, Li P, Cui S S, Cui D X, Jin H. Metal-organic frame-work engineered corn-like SERS active Ag@Carbon with controlla-ble spacing distance for tracking trace amount of organic com-pounds[J]. J. Hazard. Mater., 2022, 424:  127686. doi: 10.1016/j.jhazmat.2021.127686

    21. [21]

      Xu Y J, Shi L X, Jing X H, Miao H Y, Zhang Y. SERS-active com-posites with Au-Ag Janus nanoparticles/perovskite in immunoassays for staphylococcus aureus enterotoxins[J]. ACS App.l Mater. Interfaces, 2022, 14(2):  3293-3301. doi: 10.1021/acsami.1c21063

    22. [22]

      Sun H Z, Cong S, Zheng Z H, Wang Z, Chen Z G, Zhao Z G. Metal-organic frameworks as surface enhanced Raman scattering sub-strates with high tailorability[J]. J. Am. Chem. Soc., 2018, 141(2):  870-878.

    23. [23]

      Sun H Z, Song G, Lu W B, Cong S, Zhao Z G, Gong W. Stabilizing photo-induced vacancy defects in MOF matrix for high-performance SERS detection[J]. Nano Res, 2022, 15(6):  5347-5354. doi: 10.1007/s12274-022-4185-x

    24. [24]

      Osterrieth J W M, Wright D, Noh H, Kung C W, Vulpe D, Li A, Park J E, Jimenez D F. Core-shell gold nanorod@zirconium-based metal-organic framework composites as in situ size-selective Raman probes[J]. J. Am. Chem. Soc., 2019, 141(9):  3893-3900. doi: 10.1021/jacs.8b11300

    25. [25]

      Li J, Liu Z F, Tian D H, Li B J, Shao L, Lou Z Z. Assembly of gold nanorods functionalized by zirconium-based metal-organic frame-works for surface enhanced Raman scattering[J]. Nanoscale, 2022, 14(14):  5561-5568. doi: 10.1039/D2NR00298A

    26. [26]

      郑龙珍, 亢晓卫, 纪忆, 邹志君, 王益民, 陈计芳. Ag/ZIF-90自组装薄膜的制备及SERS活性研究[J]. 无机化学学报, 2015,31,(3): 465-471. ZHENG L Z, KANG X W, JI Y, ZOU Z J, WANG Y M, CHEN J F. Preparation of Ag/ZIF-90 self-assembled membrane and its high SERS performance[J]. Chinese J. Inorg. Chem., 2015, 31(3):  465-471.

    27. [27]

      Sun Y, Yu X X, Hu J Y, Zhuang X M, Wang J J, Qiu H X, Ren H T, Zhang S H, Zhang Y S, Hu Y J. Constructing a highly sensitivity SERS sensor based on a magnetic metal-organic framework (MOF) to detect the trace of thiabendazole in fruit juice[J]. ACS Sustain Chem. Eng., 2022, 10:  8400-8410.

    28. [28]

      Liao J, Wang D M, Liu A Q, Hu Y L, Li G K. Controlled stepwise-synthesis of core-shell Au@MIL-100(Fe) nanoparticles for sensitive surface-enhanced Raman scattering detection[J]. Analyst, 2015, 140(24):  8165-8171.

    29. [29]

      Cai Y Z, Wu Y P, Xuan T, Guo X Y, Wen Y, Yang H F. Core-shell Au@metal-organic frameworks for promoting Raman detection sensi-tivity of methenamine[J]. ACS Appl. Mater. Interfaces, 2018, 10(18):  15412-15417.

    30. [30]

      He J C, Dong J W, Hu Y F, Li G K, Hu Y L. Design of Raman tag-bridged core-shell Au@Cu-3(BTC)(2) nanoparticles for Raman imag-ing and synergistic chemo-photothermal therapy[J]. Nanoscale, 2019, 11(13):  6089-6100.

    31. [31]

      Zheng G C, de Marchi S, Lopez-Puente V, Sentosun K, Sentosun L, Perez-Juste I, Hill E H, Bals S, Liz-Marzán L M, Pastoriza-Santos I, Perez-Juste J. Encapsulation of single plasmonic nanoparticles within ZIF-8 and SERS analysis of the MOF flexibility[J]. Small, 2016, 12(29):  3935-3943.

    32. [32]

      Jiang P C, Hu Y, Li G K. Biocompatible Au@Ag nanorod@ZIF-8 core-shell nanoparticles for surface-enhanced Raman scattering imaging and drug delivery[J]. Talanta, 2019, 200:  212-217.

    33. [33]

      Li D, Cao X K, Zhang Q M, Ren X G, Jiang L, Li D W, Deng W, Liu H T. Facile in situ synthesis of core-shell MOF@Ag nanoparticle composites on screen-printed electrodes for ultrasensitive SERS detection of polycyclic aromatic hydrocarbons[J]. J. Mater. Chem. A, 2019, 7(23):  14108-14117.

    34. [34]

      Zhang P, Liu G Q, Feng S J, Zhou X, Xu W S, Cai W P. Engineering of flexible granular Au nanocap ordered array and its surface en-hanced Raman spectroscopy effect[J]. Nanotechnology, 2020, 31(3):  035303.

  • Figure 1  Fabrication process of MIL-101/Au (a); SEM images of MIL-101 (b) and MIL-101/Au (c); TEM images of MIL-101 (d) and MIL-101/Au (e); FTIR spectra (f) and XRD patterns (g) of MIL-101 and MIL-101/Au

    Figure 2  SEM images of the composite substrates of MIL-101/Au with volume ratios of 1∶1 (a), 1∶2 (b), and 1∶3 (c)

    Figure 3  SERS spectra of 10-6 mol·L-1 R6G molecules at different volume ratios of MIL-101 dispersion liquid and HAuCl 4 solution as SERS substrates (a) and the corresponding Raman intensity at 1 506 cm-1 (b); TG curves of three composite substrates and pure MIL-101 (c)

    Figure 4  SERS spectra of 10-6 mol·L-1 R6G molecules with MIL-101/Au at different adsorption times (a) and the corresponding Raman intensity at 1 506 cm-1 (b)

    Figure 5  SERS spectra of R6G with different concentrations on the MIL-101/Au substrates (a) and calibration plot based on Raman intensity at 1 506 cm-1 (b)

    Figure 6  SERS spectra of 10-6 mol·L-1 R6G collected from 20 different points on the MIL-101/Au substrates (a) and the corresponding Raman intensity distribution at 1 506 cm-1 (b)

    Figure 7  SERS spectra of 10-6 mol·L-1 R6G with MIL-101/Au obtained during 1-5 detection-washing cycles

    Figure 8  SERS spectra of thiram with different concentrations on the MIL-101/Au substrates (a) and calibration plot based on Raman intensity at 1 378 cm-1 (b)

  • 加载中
计量
  • PDF下载量:  0
  • 文章访问数:  20
  • HTML全文浏览量:  2
文章相关
  • 发布日期:  2024-10-10
  • 收稿日期:  2024-02-03
  • 修回日期:  2024-07-16
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