Next Article in Journal
ZrP2O7 as a Cathodic Material in Single-Chamber MFC for Bioenergy Production
Next Article in Special Issue
The Influence of Medium on Fluorescence Quenching of Colloidal Solutions of the Nd3+: LaF3 Nanoparticles Prepared with HTMW Treatment
Previous Article in Journal
Stainless Steel-Supported Amorphous Nickel Phosphide/Nickel as an Electrocatalyst for Hydrogen Evolution Reaction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

One-Step Hydrothermal Synthesis of Highly Fluorescent MoS2 Quantum Dots for Lead Ion Detection in Aqueous Solutions

School of Physics and Electronic Engineering, Henan Key Laboratory of Magnetoelectronic Information Functional Materials, Zhengzhou University of Light Industry, Zhengzhou 450002, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2022, 12(19), 3329; https://doi.org/10.3390/nano12193329
Submission received: 17 August 2022 / Revised: 12 September 2022 / Accepted: 21 September 2022 / Published: 24 September 2022
(This article belongs to the Special Issue Fluorescent Quantum Dot Nanomaterials)

Abstract

:
Lead ions in water are harmful to human health and ecosystems because of their high toxicity and nondegradability. It is important to explore effective fluorescence probes for Pb2+ detection. In this work, surface-functionalized molybdenum disulfide quantum dots (MoS2 QDs) were prepared using a hydrothermal method, and ammonium tetrathiomolybdate and glutathione were used as precursors. The photoluminescence quantum yield of MoS2 QDs can be improved to 20.4%, which is higher than that for MoS2 QDs reported in current research. The as-prepared MoS2 QDs demonstrate high selectivity and sensitivity for Pb2+ ions, and the limit of detection is 0.056 μM. The photoluminescence decay dynamics for MoS2 QDs in the presence of Pb2+ ions in different concentrations indicate that the fluorescence quenching originated from nonradiative electron transfer from excited MoS2 QDs to the Pb2+ ion. The prepared MoS2 QDs have great prospect and are expected to become a good method for lead ion detection.

1. Introduction

Heavy metal contamination in the environment has become an urgent problem to solve because of its threat to human health and ecosystems [1,2]. In industry and agricultural fields, heavy metal ions (Pb2+, Hg2+, Cd2+, etc.) are widely used. If not handled properly, these ions can leak into water circulation systems and further contaminate our drinking water and food, ultimately enriching in the human body and changing protein structure to cause a series of diseases. Lead pollutants are considered to be one of the most dangerous contaminants, which exhibit high toxicity and nondegradability [3,4]. So far, atomic absorption spectrometry, electrochemical technique, inductively coupled plasma mass spectrometry, among others, have been developed to detect heavy metal ions [5,6,7]. However, their application is often limited due to complicated sample pretreatment, long analysis time, and expensive equipment. To overcome these shortcomings, an approach of optical detection based on fluorescence analysis has emerged for its high sensitivity and selectivity. Hence, to detect Pb2+ ions, it is urgent to develop an ecofriendly fluorescent material with high sensitivity and selectivity.
Owing to quantum confinement effect and edge effect, semiconductor quantum dots (QDs) possess unique photophysical properties and can be used as a fluorescence probe for ion detection. Fluorescence probes for Pb2+ based on QDs have gained much interest. Mn-doped ZnS QDs, ZnSeS/Cu:ZnS/ZnS core/shell/shell QDs, and carbon dots have been developed for Pb2+ detection [8,9,10]. However, the inorganic quantum dots are imperfect for their toxicity and multistep synthetic approach. Although carbon dots have created great focus for their facile synthesis and low toxicity, low quantum yield limits their application. In recent years, two-dimensional transition metal dichalcogenides such as MoS2 and WSe2, and especially their quantum dots, have received much attention in sensing applications and optoelectronic devices [11,12,13,14,15,16,17]. MoS2 QDs with large surface-to-volume ratio and abundant active edge sites can be used as a photoluminescence-sensing platform [18,19,20,21]. Different from traditional quantum dots (CdSe, CdTe, etc.) with harmful elements, MoS2 QDs are water-soluble and nontoxic. Hence, MoS2 QDs have received extensive attention in bioimaging, sensing and photodynamic therapy [22,23]. MoS2 QDs have been used as a sensor to detect nitro explosives, hyaluronidase and hydrogen peroxide, along with glucose and other biomolecules [18,19,20,24,25,26]. Additionally, there are some reports on the metal ion sensor using fluorescent MoS2 QDs. Cysteine-functionalized MoS2 QDs have been used for sensing Al3+ and Fe3+ metal ions [27]. Based on the fluorescence turn-off effect, MoS2 QDs can be used as a sensor for Fe3+ detection [28]. However, the fluorescence quantum yield of these MoS2 QDs reported by previous research is less than 10%, and the interaction of metal ions with MoS2 QDs is not well studied. To obtain a high fluorescence MoS2 QDs, a widely used bottom-up method of hydrothermal synthesis has been improved with different molybdenum and sulfur sources, which is simple, ecofriendly and easy to operate. In the hydrothermal method, MoS2 QDs were obtained by the Xian group; they successfully adopted ammonium tetrathiomolybdate [(NH4)2MoS4] and hydrazine hydrate as precursor and reducing agent, respectively [24]. Zhang et al. have also synthesized MoS2 QDs using this method, with sodium molybdate (Na2MoO4·2H2O) and glutathione (GSH) serving as molybdenum and sulfur sources [27]. Although the photoluminescence quantum yield of MoS2 QDs can be increased to 6% in this method, it is challenging to produce large quantities of bright MoS2 QDs by using more appropriate Mo and S sources.
Herein, we explore high fluorescent MoS2 QDs in a one-step hydrothermal method using (NH4)2MoS4 and GSH as Mo and S sources. On the one hand, GSH was used as a reductant to reduce (NH4)2MoS4; on the other hand, GSH, as a passivation agent, could eliminate the surface defects of MoS2 QDs to enhance the fluorescence. If we introduced the Pb2+ ions into the resultant MoS2 QDs aqueous solution, quenched fluorescence was observed. We further acquired the linear dependence of fluorescence intensity on Pb2+ ion concentration in a certain range. The limit of detection (LOD) is 0.056 μM, which is below the acceptable limit given by United States Environmental Protection Agency. We further study the photoluminescence decay dynamics of MoS2 QDs with increasing the concentration of Pb2+ ion. With a detailed exciton dynamics study, we found that the fluorescence quenching originated from the electron transfer from MoS2 QDs to Pb2+ ions. As a result, MoS2 QDs prepared in this work can be used as a sensing probe to detect Pb2+ ions in water with high sensitivity and selectivity. These results are important to understanding the sensing mechanism of MoS2 QDs to metal ions.

2. Materials and Methods

2.1. Chemicals

All chemicals were of analytical grade without further purification. (NH4)2MoS4 (J&K Chemical Ltd., Shanghai, China) and GSH (Aladdin Industrial Co., Ltd., Shanghai, China) were used as precursor and reductant respectively. The metal analytes used in this study were CaCl2, MgCl2·6H2O, MnCl2·4H2O, PbCl2, AlCl3, SnCl2, FeCl2·4H2O, CuCl2, Fe(NO3)3·9H2O, HgCl2, BaBr2, ZnCl2, CsCl, KCl and AgNO3. In addition, HCl (0.1 M) and NaOH (0.1 M) solution was prepared to adjust pH. Quinine sulfate (0.05 M) was chosen as a standard sample to calculate the photoluminescence quantum yield of MoS2 QDs in a reference method. Ultrapure water was used throughout the experiment.

2.2. Preparation of MoS2 QDs

(NH4)2MoS4 and GSH were chosen as Mo and S sources to synthesis MoS2 QDs through a pot hydrothermal process. Briefly, (NH4)2MoS4 (65 mg) was dispersed in ultrapure water (25 mL), sonicated for 10 min until the (NH4)2MoS4 was fully dissolved. Afterward, 0.1 M HCl was used to adjust the pH of solution to 6.5. In addition, GSH (308 mg) was dissolved in ultrapure water (50 mL), and subsequently added into the (NH4)2MoS4 aqueous solution prepared in the previous step. The mixture was further sonicated for 10 min and heated at 200 °C in a Teflon-lined autoclave (100 mL) for 24 h. The solution was then naturally cooled to room temperature. The mixture was centrifuged for 10 min at 10,000 rpm to collect the suspension, subsequently purified by a 0.22 µm microporous membrane. The target MoS2 QDs aqueous solution was stored at 4 °C for further characterization.

2.3. Characterization of MoS2 QDs

Commercial experimental instruments including a UV–Vis–NIR spectrophotometer (Cary 5000, Agilent, Santa Clara, CA, USA) and fluorescence spectrometer (PerkinElmer LS 55, Waltham, MA, USA) were used to measure the absorption and photoluminescence spectra, respectively. Time-resolved photoluminescence (TRPL) measurements were conducted with a home-built fluorescence lifetime setup (Picoharp300, Picoquant, Berlin, Germany) based on a time-correlated single photon counting (TCSPC) module. A pulsed excitation source (350 nm with repetition frequency of 13 MHz) was used to excite the samples.
The particle diameters and monodispersity of MoS2 QDs were characterized by high resolution transmission electron microscopy (HRTEM) (JEM-2010, JEOL, Tokyo, Japan). The aqueous solution of appropriate concentration was dripped onto a carbon-coated copper grid, and the aqueous solvent was dried by nitrogen flow at ambient temperature.
The monolayer nature of MoS2 QDs was checked by atomic force microscopy (AFM, Solver-P47H, NT-MDT, Moscow, Russia). The sample for AFM measurement was acquired by spin coating diluted MoS2 aqueous solution on a mica wafer.
Elemental analysis was conducted using X-ray photoelectron spectroscopy (XPS) (PHI 5000 VersaProbeIII, Japan). The binding energy calibration was performed using C 1s of 284.6 eV as standard peak energy.
The surface functional groups of MoS2 QDs were identified by Fourier transform infrared spectroscopy (FTIR) spectra (Perkin-Elmer spectrometer, Spectrum One B, Waltham, MA, USA), and the MoS2 QDs powder was pressed into a tablet with KBr.

2.4. Procedures for Detection of Metal Ions

For the selectivity study of MoS2 QDs, an aqueous solution of metal ions including Ca2+, Mg2+, Mn2+, Pb2+, Al3+, Sn2+, Fe2+, Cu2+, Fe3+, Hg2+, Ba2+, Zn2+, Cs2+, K+ and Ag+ was prepared with concentration of 25 μM. For the sensitivity study, Pb2+ ion solution was prepared with concentrations of 120 μM, 100 μM, 90 μM, 80 μM, 70 μM, 60 μM, 50 μM, 45 μM, 40 μM, 35 μM, 30 μM, 25 μM, 20 μM, 15 μM, 10 μM and 5 μM. Then, the various metal ions were added into the corresponding MoS2 QDs solution to measure the fluorescent intensity by using the fluorescence spectrometer.

3. Results

The MoS2 QDs were first characterized by TEM. The MoS2 QDs are monodispersed and homogeneous with an average diameter of 4.4 ± 0.2 nm (Figure 1a,b). The inset in Figure 1a is the HRTEM image of a single MoS2 QD with ordered lattice fringes. The lattice spacing is ~2.3 Å, which can be attributed to the (103) plane of crystalline MoS2. AFM was further used to check the thickness of the MoS2 QDs. Figure 1c also shows the monodispersity and size uniformity of the MoS2 QDs. Additionally, the profile shown in Figure 1d indicates that the height of MoS2 QDs is ~0.7 nm, demonstrating a monolayer nature [29]. The optical properties of absorption and photoluminescence spectra are presented in Figure 1e,f. A typical excitonic peak of MoS2 QDs at 310 nm with a remarkable shoulder is shown in Figure 1e [22]. The fluorescence emission peak is 430 nm under 350 nm excitation (Figure 1f). The photoluminescence quantum yield of MoS2 QDs was measured choosing quinine sulfate (54%, 350 nm excitation) as a standard sample. According to the data in Figure 1e,f, the photoluminescence quantum yield of MoS2 QDs is calculated at 20.4%, which is relatively higher than those reported previously for MoS2 QDs fabricated with a similar hydrothermal method [24,27]. In this study, the GSH was used as a reductant and passivation agent, which can provide enough surface functional groups to eliminate the edge defects, resulting in a high fluorescence emission.
The high-resolution XPS was conducted for further elementary analysis. The XPS response in the Mo 3d, S 2p is shown in Figure 2a,b, respectively. Figure 2a exhibits two peaks at 232.8 and 229.5 eV, which correspond to Mo4+ 3d3/2 and Mo4+ 3d5/2. Additionally, the two characteristic peaks of S 2p1/2 and S 2p3/2 located at 163.4 and 162.3 eV indicate a 2H phase for the crystal structure of MoS2 QDs [30,31]. Moreover, the atomic ratio of Mo/S is about 1:2, indicating the formation of MoS2 QDs. The formula of GSH is drawn in Figure 2c. During synthesis of MoS2 QDs, GSH was not only used as the reducing agent, but also as surface passivation agent to provide carboxyl, amino, and thiol groups for MoS2 QDs. The surface functional groups of MoS2 QDs were further verified by FTIR (Figure 2d). The characteristic peaks at 3212 and 3100 cm−1 (N-H stretching vibrations), 2600 cm−1 (S-H stretching vibration), 1684 cm−1 (C=O stretching vibration), 1590 cm−1 (N-H in-plane bending vibration), 1404 cm−1 (C-N stretching vibration of amide group), 1280 cm−1 (C-N stretching vibration of amine group), 1235 cm−1 (O-H stretching vibration), 1109 cm−1 (C-O stretching vibration) and 740 cm−1 (N-H out-plane bending vibration) indicate that carboxyl, amino and thiol groups of GSH can decorate the surface of MoS2 QDs to eliminate the edge defects, resulting in a high fluorescence emission.
The high fluorescence and environment-friendly characteristics of surface-functionalized MoS2 QDs make them a potential candidate for sensing metal ions. To explore the sensing capability of these MoS2 QDs, selectivity toward various metal ions was performed. Various metal ions with the same concentration of 25 μM were added into the MoS2 QDs aqueous solution to study the influence on respective MoS2 QDs fluorescence. The (F0 − F)/F0 value was used to determine the fluorescence enhancement or quenching of MoS2 QDs, where the fluorescence intensities of MoS2 QDs without/with metal ions were represented by F0 and F, respectively. Figure 3a indicates that only Pb2+ ions caused obvious fluorescence quenching, while the other metal ions have a slight impact on the fluorescence intensity. These results demonstrate that the surface-functionalized MoS2 QDs have high selectivity for Pb2+ ion.
To evaluate the stability of MoS2 QDs under different conditions, the effect of pH on the fluorescence of MoS2 QDs was also studied by adjusting the acidity to alkalinity. As shown in Figure 3b, the fluorescence intensity of MoS2 QDs shows a slight variation with increasing the pH in a wide range from 3.7 to 10.4, indicating negligible influence of pH on the fluorescence of MoS2 QDs. When the sample was kept for 50 days, the fluorescence intensity of MoS2 QDs was almost unchanged while the position of maximum emission peak was red shifted slightly (Figure 3c). These observations suggest that MoS2 QDs can be used as a stable fluorescence probe in complex underwater environments.
The sensitivity for Pb2+ ion detection was carried out by fluorescence titration of MoS2 QDs with varying concentration of Pb2+ ion from 0 to 120 μM. By increasing the concentration of Pb2+ ions from 0 to 120 μM, the fluorescence intensity of the MoS2 QDs decreases gradually (Figure 4a). Additionally, we plotted the degree of quenching (F0 − F)/F0 versus Pb2+ ion concentration in the range 0 to 60 μM (Figure 4b). We obtained a good linear relationship between (F0 − F)/F0 with Pb2+ ion concentration (R2 = 0.984), and the LOD was measured at 0.056 μM (3σ per slope, where σ is the standard deviation of blank signals, N = 10), which is less than the 15 μg/L safety value set by United States Environmental Protection Agency [32]. Moreover, compared with carbon dots and other inorganic probes used for the detection of Pb2+ by using fluorescence method (Table 1), the LOD obtained using MoS2 QDs as the sensor was slightly larger while the linear detection range was wider. However, compared with 1,4-diaminobutane (DAB) capped MoS2 QDs for monitoring the Pb2+ ions [21], the GSH-functionalized MoS2 QDs in this work possess higher quantum yield and lower LOD, which is suitable to use as a fluorescent probe for lead ions. These results suggest that the surface-functionalized MoS2 QDs can be used as a sensor for Pb2+ ion detection with high selectivity and sensitivity.
To further study the fluorescence quenching mechanism of MoS2 QDs for Pb2+ ion detection, we measured the photoluminescence decay dynamics of MoS2 QDs in the presence of Pb2+ ion (0, 5, 10, 15, 20 μM). The photoluminescence lifetime of MoS2 QDs decreases gradually with increasing concentration of Pb2+ ion (Figure 5a), indicating that Pb2+ ion plays an important role in the exciton recombination of MoS2 QDs. The photoluminescence decay curves in Figure 5a can be fitted with a bi-exponential model; the fitting parameters are shown in Table 2. There are two decay processes (short lifetime τ1 and long lifetime τ2) for the exciton deactivation. The fast component τ1 shows a mild variation while slow component τ2 becomes shorter with increasing concentration of Pb2+ ions. These two lifetime components could be assigned to band-edge emission and surface-state-assisted emission, respectively [21,36]. The shorter of slow component τ2 from surface-state-assisted emission may be ascribed to the observation that Pb2+ ion can chelate with surface functional groups of MoS2 QDs [33], resulting in electron transfer from the excited-state MoS2 QDs to the Pb2+ ion. The electron transfer rate κET can be calculated by the following equation [37]:
κ E T = 1 τ av 1 τ 0 ,  
where τav and τ0 are the average lifetime of MoS2 QDs with and without Pb2+ ions, respectively. When the Pb2+ ion concentration is 5 μM, the κET is 3.26 × 107 s−1. We further observed the linear dependence of electron transfer rate κET on the Pb2+ ion concentration (Figure 5b). This evolution of κET further indicates that the fluorescence quenching originates from nonradiative electron transfer from the excited MoS2 QDs to the Pb2+ ions.

4. Conclusions

We prepared high fluorescence MoS2 QDs by using an easily manipulated hydrothermal method, in which (NH4)2MoS4 and GSH were used as Mo and S sources. In this method, GSH was used as a reductant and capping agent to obtain high fluorescent MoS2 QDs. The morphological and structural characterization from TEM and AFM demonstrated that the ultrasmall MoS2 QDs (~4.4 nm) were monodispersed and single-layered. FTIR analysis further manifested that the functional groups provided by GSH can passivate the surface of MoS2 QDs. The photoluminescence of MoS2 QDs was quenched in the presence of Pb2+ ions. Based on the quenching effect, the surface-functionalized MoS2 QDs were used as a fluorescence probe with high sensitivity and selectivity to detect Pb2+ ions. From the fluorescence titration of MoS2 QDs, we determined the sensitivity of this sensor at the detection limit of 0.056 μM. Moreover, the photoluminescence decay dynamics of MoS2 QDs help us to understand the electron transfer behavior between MoS2 QDs and Pb2+ ions. A linear relationship between electron transfer rate κET and Pb2+ ion concentration was observed. Our findings indicate that this water-soluble nanomaterial holds great promise and would be a good sensor for the detection of lead ions.

Author Contributions

Conceptualization, D.L.; methodology S.F.; investigation and data curation, L.X.; writing—original draft preparation, L.X.; writing—review and editing, L.X.; supervision, G.G.; funding acquisition, Y.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially sponsored by the Emergency Management Project of National Natural Science Foundation of China (Project No. 11647084), the Key Scientific Research Project of Colleges and Universities in Henan Province (Project No. 21B140012) and the Key Research & Development Promotion Projects in Henan Province (Project No. 212102210132).

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Guidi, P.; Bernardeschi, M.; Palumbo, M.; Genovese, M.; Scarcelli, V.; Fiorati, A.; Riva, L.; Punta, C.; Corsi, I.; Frenzilli, G. Suitability of a cellulose-based nanomaterial for the remediation of heavy metal contaminated freshwaters: A case-study showing the recovery of cadmium induced DNA integrity loss, cell proliferation increase, nuclear morphology and chromosomal alterations on dreissena polymorpha. Nanomaterials 2020, 10, 1837. [Google Scholar]
  2. Zhang, Y.; Wang, B.; Cheng, Q.; Li, X.; Li, Z. Removal of toxic heavy metal ions (Pb, Cr, Cu, Ni, Zn, Co, Hg, and Cd) from waste batteries or lithium cells using nanosized metal oxides: A review. J. Nanosci. Nanothechnol. 2020, 20, 7231–7254. [Google Scholar] [CrossRef] [PubMed]
  3. Jeevanaraj, P.; Ahmad Foat, A.; Tholib, H.; Ahmad, N.I. Heavy metal contamination in processed seafood and the associated health risk for malaysian women. Br. Food J. 2020, 122, 3099–3114. [Google Scholar] [CrossRef]
  4. Singh, H.; Bamrah, A.; Bhardwaj, S.K.; Deep, A.; Khatri, M.; Kim, K.H.; Bhardwaj, N. Nanomaterial-based fluorescent sensors for the detection of lead ions. J. Hazard. Mater. 2021, 407, 124379. [Google Scholar] [CrossRef]
  5. Balcaen, L.; Bolea-Fernandez, E.; Resano, M.; Vanhaecke, F. Inductively coupled plasma-tandem mass spectrometry (ICP-MS/MS): A powerful and universal tool for the interference-free determination of (ultra)trace elements—A tutorial review. Anal. Chim. Acta 2015, 894, 7–19. [Google Scholar] [CrossRef] [PubMed]
  6. Niu, X.; Zhong, Y.; Chen, R.; Wang, F.; Liu, Y.; Luo, D. A “turn-on” fluorescence sensor for Pb2+ detection based on graphene quantum dots and gold nanoparticles. Sens. Actuators B Chem. 2018, 255, 1577–1581. [Google Scholar] [CrossRef]
  7. Shi, J.J.; Zhu, J.C.; Zhao, M.; Wang, Y.; Yang, P.; He, J. Ultrasensitive photoelectrochemical aptasensor for lead ion detection based on sensitization effect of CdTe QDs on MoS2-CdS: Mn nanocomposites by the formation of G-quadruplex structure. Talanta 2018, 183, 237–244. [Google Scholar] [CrossRef]
  8. Chen, J.; Zhu, Y.; Zhang, Y. Glutathione-capped Mn-doped ZnS quantum dots as a room-temperature phosphorescence sensor for the detection of Pb2+ Ions. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2016, 164, 98–102. [Google Scholar] [CrossRef]
  9. Mabrouk, S.; Rinnert, H.; Balan, L.; Jasniewski, J.; Medjahdi, G.; Ben Chaabane, R.; Schneider, R. Aqueous synthesis of core/shell/shell ZnSeS/Cu:ZnS/ZnS quantum dots and their use as a probe for the selective photoluminescent detection of Pb2+ in water. J. Photochem. Photobiol. A Chem. 2022, 431, 114050. [Google Scholar] [CrossRef]
  10. Kumar, A.; Chowdhuri, A.R.; Laha, D.; Mahto, T.K.; Karmakar, P.; Sahu, S.K. Green synthesis of carbon dots from ocimum sanctum for effective fluorescent sensing of Pb2+ ions and live cell imaging. Sens. Actuators B Chem. 2017, 242, 679–686. [Google Scholar] [CrossRef]
  11. Dhenadhayalan, N.; Lin, T.W.; Lee, H.L.; Lin, K.C. Multisensing capability of MoSe2 quantum dots by tuning surface functional groups. ACS Appl. Nano Mater. 2018, 1, 3453–3463. [Google Scholar] [CrossRef]
  12. Lin, T.W.; Dhenadhayalan, N.; Lee, H.L.; Lin, Y.T.; Lin, K.C.; Chang, A. Fluorescence turn-on chemosensors based on surface-functionalized MoS2 quantum dots. Sens. Actuators B Chem. 2019, 281, 659–669. [Google Scholar] [CrossRef]
  13. Zhang, Y.; Zhou, S.; Liu, H.; Tang, X.; Zhou, H.; Cai, H. Nitrogen-doped MoS2 QDs as fluorescent probes for sequential dual-target detection and their microfluidic logic gate operations. Microchem. J. 2021, 169, 106553. [Google Scholar] [CrossRef]
  14. Lan, Y.W.; Torres, C.M., Jr.; Tsai, S.H.; Zhu, X.; Shi, Y.; Li, M.Y.; Li, L.J.; Yeh, W.K.; Wang, K.L. Atomic-monolayer MoS2 band-to-band tunneling field-effect transistor. Small 2016, 12, 5676–5683. [Google Scholar] [CrossRef]
  15. Szczęśniak, D.; Hoehn, R.D.; Kais, S. Canonical schottky barrier heights of transition metal dichalcogenide monolayers in contact with a metal. Phys. Rev. B 2018, 97, 195315. [Google Scholar] [CrossRef]
  16. Pak, J.; Lee, I.; Cho, K.; Kim, J.-K.; Jeong, H.; Hwang, W.-T.; Ahn, G.H.; Kang, K.; Yu, W.J.; Javey, A. Intrinsic optoelectronic characteristics of MoS2 phototransistors via a fully transparent van der Waals heterostructure. ACS Nano 2019, 13, 9638–9646. [Google Scholar] [CrossRef] [PubMed]
  17. Szczęśniak, D.; Kais, S. Gap states and valley-spin filtering in transition metal dichalcogenide monolayers. Phys. Rev. B 2020, 101, 115423. [Google Scholar] [CrossRef]
  18. Wang, Y.; Ni, Y. Molybdenum disulfide quantum dots as a photoluminescence sensing platform for 2, 4, 6-trinitrophenol detection. Anal. Chem. 2014, 86, 7463–7470. [Google Scholar] [CrossRef]
  19. Haldar, D.; Dinda, D.; Saha, S.K. High selectivity in water soluble MoS2 quantum dots for sensing nitro explosives. J. Mater. Chem. C 2016, 4, 6321–6326. [Google Scholar] [CrossRef]
  20. Wang, X.; Wu, Q.; Jiang, K.; Wang, C.; Zhang, C. One-step synthesis of water-soluble and highly fluorescent MoS2 quantum dots for detection of hydrogen peroxide and glucose. Sens. Actuators B Chem. 2017, 252, 183–190. [Google Scholar] [CrossRef]
  21. Sharma, P.; Mehata, M.S. Rapid sensing of lead metal ions in an aqueous medium by MoS2 quantum dots fluorescence turn-off. Mater. Res. Bull. 2020, 131, 110978. [Google Scholar] [CrossRef]
  22. Dong, H.; Tang, S.; Hao, Y.; Yu, H.; Dai, W.; Zhao, G.; Cao, Y.; Lu, H.; Zhang, X.; Ju, H. Fluorescent MoS2 quantum dots: Ultrasonic preparation, up-conversion and down-conversion bioimaging, and photodynamic therapy. ACS Appl. Mater. Interfaces 2016, 8, 3107–3114. [Google Scholar] [CrossRef] [PubMed]
  23. Xu, S.; Li, D.; Wu, P. One-pot, facile, and versatile synthesis of monolayer MoS2/WS2 quantum dots as bioimaging probes and efficient electrocatalysts for hydrogen evolution reaction. Adv. Funct. Mater. 2015, 25, 1127–1136. [Google Scholar] [CrossRef]
  24. Gu, W.; Yan, Y.; Cao, X.; Zhang, C.; Ding, C.; Xian, Y. A facile and one-step ethanol-thermal synthesis of MoS2 quantum dots for two-photon fluorescence imaging. J. Mater. Chem. B 2016, 4, 27–31. [Google Scholar] [CrossRef] [PubMed]
  25. Swaminathan, H.; Balasubramanian, K. Förster resonance energy transfer between MoS2 quantum dots and polyaniline for turn-on bovine serum albumin sensing. Sens. Actuators B Chem. 2018, 264, 337–343. [Google Scholar] [CrossRef]
  26. Yang, L.; Ge, J.; Ma, D.; Tang, J.; Wang, H.; Li, Z. MoS2 quantum dots as fluorescent probe for methotrexate detection. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2022, 279, 121443. [Google Scholar] [CrossRef] [PubMed]
  27. Wu, Q.; Wang, X.; Jiang, Y.; Sun, W.; Wang, C.; Yang, M.; Zhang, C. MoS2-QD-based dual-model photoluminescence sensing platform for effective determination of Al3+ and Fe3+ simultaneously in various environment. ChemistrySelect 2018, 3, 2326–2331. [Google Scholar] [CrossRef]
  28. Ma, J.; Yu, H.; Jiang, X.; Luo, Z.; Zheng, Y. High sensitivity label-free detection of Fe3+ ion in aqueous solution using fluorescent MoS2 quantum dots. Sens. Actuators B Chem. 2019, 281, 989–997. [Google Scholar] [CrossRef]
  29. Ha, H.D.; Han, D.J.; Choi, J.S.; Park, M.; Seo, T.S. Dual role of blue luminescent MoS2 quantum dots in fluorescence resonance energy transfer phenomenon. Small 2014, 10, 3858–3862. [Google Scholar] [CrossRef] [PubMed]
  30. Bang, G.S.; Nam, K.W.; Kim, J.Y.; Shin, J.; Choi, J.W.; Choi, S.Y. Effective liquid-phase exfoliation and sodium ion battery application of MoS2 nanosheets. ACS Appl. Mater. Interfaces 2014, 6, 7084–7089. [Google Scholar] [CrossRef] [PubMed]
  31. Pagona, G.; Bittencourt, C.; Arenal, R.; Tagmatarchis, N. Exfoliated semiconducting pure 2H-MoS2 and 2H-WS2 assisted by chlorosulfonic acid. Chem. Commun. 2015, 51, 12950–12953. [Google Scholar] [CrossRef] [PubMed]
  32. Kaewprom, C.; Sricharoen, P.; Limchoowong, N.; Nuengmatcha, P.; Chanthai, S. Resonance light scattering sensor of the metal complex nanoparticles using diethyl dithiocarbamate doped graphene quantum dots for highly Pb (II)-sensitive detection in water sample. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2019, 207, 79–87. [Google Scholar] [CrossRef] [PubMed]
  33. Liu, Y.; Zhou, Q.; Li, J.; Lei, M.; Yan, X. Selective and sensitive chemosensor for lead ions using fluorescent carbon dots prepared from chocolate by one-step hydrothermal method. Sens. Actuators B Chem. 2016, 237, 597–604. [Google Scholar] [CrossRef]
  34. Wang, X.Y.; Niu, C.G.; Guo, L.J.; Hu, L.Y.; Wu, S.Q.; Zeng, G.M.; Li, F. A fluorescence sensor for lead (II) ions determination based on label-free gold nanoparticles (GNPs)-DNAzyme using time-gated mode in aqueous solution. J. Fluoresc. 2017, 27, 643–649. [Google Scholar] [CrossRef] [PubMed]
  35. Wang, X.; Guo, H.; Wu, N.; Xu, M.; Zhang, L.; Yang, W. A dual-emission fluorescence sensor constructed by encapsulating double carbon dots in zeolite imidazole frameworks for sensing Pb2+. Colloids Surf. Physicochem. Eng. Asp. 2021, 615, 126218. [Google Scholar] [CrossRef]
  36. Shi, H.; Yan, R.; Bertolazzi, S.; Brivio, J.; Gao, B.; Kis, A.; Jena, D.; Xing, H.G.; Huang, L. Exciton dynamics in suspended monolayer and few-layer MoS2 2D crystals. ACS Nano 2013, 7, 1072–1080. [Google Scholar] [CrossRef]
  37. Brown, K.A.; Song, Q.; Mulder, D.W.; King, P.W. Diameter dependent electron transfer kinetics in semiconductor–enzyme complexes. ACS Nano 2014, 8, 10790–10798. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) TEM image of monodispersed MoS2 QDs, and the inset shows the HRTEM image with lattice spacing; (b) diameter distribution histogram of MoS2 QDs; (c) AFM image of MoS2 QDs; (d) corresponding height profile of blue line labeled in (c); (e) UV–vis absorption; (f) photoluminescence spectra of MoS2 QDs.
Figure 1. (a) TEM image of monodispersed MoS2 QDs, and the inset shows the HRTEM image with lattice spacing; (b) diameter distribution histogram of MoS2 QDs; (c) AFM image of MoS2 QDs; (d) corresponding height profile of blue line labeled in (c); (e) UV–vis absorption; (f) photoluminescence spectra of MoS2 QDs.
Nanomaterials 12 03329 g001
Figure 2. (a,b) Typical XPS response in Mo 3d and S 2p regions, and the solid line shows the fitting result of experimental data; (c) formula of GSH; (d) FTIR spectrum of surface functionalized MoS2 QDs with carboxyl, amino and thiol groups.
Figure 2. (a,b) Typical XPS response in Mo 3d and S 2p regions, and the solid line shows the fitting result of experimental data; (c) formula of GSH; (d) FTIR spectrum of surface functionalized MoS2 QDs with carboxyl, amino and thiol groups.
Nanomaterials 12 03329 g002
Figure 3. (a) The fluorescence enhancement or quenching of MoS2 QDs in the presence of various metal ions. (b) Photoluminescence spectra of MoS2 QDs under different pH conditions. (c) Photoluminescence spectra of MoS2 QDs stored at 4 °C for 0 and 50 days.
Figure 3. (a) The fluorescence enhancement or quenching of MoS2 QDs in the presence of various metal ions. (b) Photoluminescence spectra of MoS2 QDs under different pH conditions. (c) Photoluminescence spectra of MoS2 QDs stored at 4 °C for 0 and 50 days.
Nanomaterials 12 03329 g003
Figure 4. (a) Fluorescence spectra of MoS2 QDs with different concentrations of Pb2+ ion from 0 to 120 μM; (b) linear curve between (F0 − F)/F0 and Pb2+ ion concentration in the range 5 to 60 μM.
Figure 4. (a) Fluorescence spectra of MoS2 QDs with different concentrations of Pb2+ ion from 0 to 120 μM; (b) linear curve between (F0 − F)/F0 and Pb2+ ion concentration in the range 5 to 60 μM.
Nanomaterials 12 03329 g004
Figure 5. (a) Photoluminescence decay curves of MoS2 QDs in presence of different concentrations of Pb2+ ions (0, 5, 10, 15, 20 μM) under pulsed excitation at 350 nm. (b) Dependence of electron transfer rate on concentrations of Pb2+ ions.
Figure 5. (a) Photoluminescence decay curves of MoS2 QDs in presence of different concentrations of Pb2+ ions (0, 5, 10, 15, 20 μM) under pulsed excitation at 350 nm. (b) Dependence of electron transfer rate on concentrations of Pb2+ ions.
Nanomaterials 12 03329 g005
Table 1. Comparison of various fluorescent sensors for Pb2+ ion detection.
Table 1. Comparison of various fluorescent sensors for Pb2+ ion detection.
SensorsLinear RangeLODReference
Carbon dots0.01–1 μM0.59 nM[10]
Carbon dots0.033–1.67 μM12.7 nM[33]
ZnSeS/Cu:ZnS/ZnS QDs0.04–6 μM21 nM[9]
Mn-doped ZnS QDs1–100 μg/L0.45 μg/L [8]
Gold nanoparticles-DNAzyme10–2500 nM1.7 nM[34]
N-CDs/R-CDs@ZIF-80.05–50 μM4.78 nM[35]
MoS2 QDS33 μM–8 mM50 μM[21]
MoS2 QDS5–60 μM 0.056 μMThis work
Table 2. Bi-exponential fitting for photoluminescence decay curves of the MoS2 QDs. Fitting equation: I(t) = I0 + A1exp(−t/τ1) + A2exp(−t/τ2), and the average lifetime τav = (A1τ12 + A2τ22)/(A1τ1 + A2τ2), where A1, A2 and τ1, τ2 are the amplitudes and lifetimes, respectively.
Table 2. Bi-exponential fitting for photoluminescence decay curves of the MoS2 QDs. Fitting equation: I(t) = I0 + A1exp(−t/τ1) + A2exp(−t/τ2), and the average lifetime τav = (A1τ12 + A2τ22)/(A1τ1 + A2τ2), where A1, A2 and τ1, τ2 are the amplitudes and lifetimes, respectively.
ConcentrationA1A2τ1 (ns)τ2 (ns)τav (ns)
0 μM0.340.153.112.89.4
5 μM0.370.093.111.67.2
10 μM0.380.092.99.55.8
15 μM0.380.082.88.65.1
20 μM0.390.082.57.54.4
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Xie, L.; Yang, Y.; Gong, G.; Feng, S.; Liu, D. One-Step Hydrothermal Synthesis of Highly Fluorescent MoS2 Quantum Dots for Lead Ion Detection in Aqueous Solutions. Nanomaterials 2022, 12, 3329. https://doi.org/10.3390/nano12193329

AMA Style

Xie L, Yang Y, Gong G, Feng S, Liu D. One-Step Hydrothermal Synthesis of Highly Fluorescent MoS2 Quantum Dots for Lead Ion Detection in Aqueous Solutions. Nanomaterials. 2022; 12(19):3329. https://doi.org/10.3390/nano12193329

Chicago/Turabian Style

Xie, Luogang, Yang Yang, Gaoshang Gong, Shiquan Feng, and Dewei Liu. 2022. "One-Step Hydrothermal Synthesis of Highly Fluorescent MoS2 Quantum Dots for Lead Ion Detection in Aqueous Solutions" Nanomaterials 12, no. 19: 3329. https://doi.org/10.3390/nano12193329

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop