Next Article in Journal
A Label-Free Optical Detection of Pathogens in Isopropanol as a First Step towards Real-Time Infection Prevention
Next Article in Special Issue
Glycated Hemoglobin and Methods for Its Point of Care Testing
Previous Article in Journal
Silicon Nanowire Field-Effect Transistor as Biosensing Platforms for Post-Translational Modification
Previous Article in Special Issue
Label-Free and Sensitive Determination of Cadmium Ions Using a Ti-Modified Co3O4-Based Electrochemical Aptasensor
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

An Electrochemical Aptasensor for Pb2+ Detection Based on Metal–Organic-Framework-Derived Hybrid Carbon

1
School of Agriculture and Biology, Shanghai Jiaotong University, Shanghai 200240, China
2
Key Laboratory of Urban Agriculture, Ministry of Agriculture and Rural Affairs of the People’s Republic of China, Shanghai Jiao Tong University, Shanghai 200240, China
*
Author to whom correspondence should be addressed.
Biosensors 2021, 11(1), 1; https://doi.org/10.3390/bios11010001
Submission received: 26 November 2020 / Revised: 14 December 2020 / Accepted: 18 December 2020 / Published: 22 December 2020
(This article belongs to the Special Issue New Developments for Efficient Rapid Bioassays)

Abstract

:
A new double-shelled carbon nanocages material was synthesized and developed an aptasensor for determining Pb2+ in aqueous solution. Herein, nanoporous carbon materials derived from core–shell zeolitic imidazolate frameworks (ZIFs) demonstrated excellent electrochemical activity, stability, and high specificity surface area, consequently resulting in the strong binding with aptamers. The aptamer strands would be induced to form G-quadruplex structure when Pb2+ was introduced. Under optimal conditions, the aptasensor exhibited a good linear relationship of Pb2+ concentration ranging from 0.1 to 10 μg L−1 with the detection limits of 0.096 μg L−1. The feasibility was proved by detecting Pb2+ in spiked water samples and polluted soil digestion solution. The proposed aptasensor showed excellent selectivity and reproducibility, indicating promising applications in environmental monitoring.

1. Introduction

Heavy metals are distributed worldwide and are harmful to human health and eco-environments [1,2,3,4]. In humans, lead is one of the most common pollutants accumulated through the food chain, thereby damaging human health even at a low concentration [2,5]. Lead ion (Pb2+) may cause a serial damaging to human health, such as memory loss, anemia, and irritability [6,7,8]. Therefore, development of new, accurate, and simple methods for Pb2+ determination is an urgent task.
In recent years, the sensors based on aptamers have shown great potential in detecting heavy metal ions. Aptamers are short and specific single-strand oligonucleotides, selected through systematic evolution of ligands by exponential enrichment (SELEX) [9,10]. The Pb2+-aptamer is a guanine (G)-rich oligonucleotides strand which could form G-quadruplex structure on interacting with Pb2+ [11,12,13]. Based on the high affinity and selectivity of aptamers, numbers of aptasensors have been developed for Pb2+ detection, such as fluorescent [14,15], colorimetric, chemiluminescent [16,17], and electrochemical analysis [18,19].
Electrochemical sensor has been identified as one of the most feasible methods for the detection of heavy metal, for the devices are not only low-cost and relatively simple but have also high selectivity and sensitivity [20,21]. Moreover, a series of nanomaterials have been used in analytical electrochemistry, including gold nanoparticles [22,23,24], metal or metal oxide nanoparticles [25,26], quantum dots [11], graphene [14,27], and metal–organic frameworks (MOFs) [28,29]. MOFs, porous inorganic–organic hybrid materials assembled from metal ions and organic ligands, have attracted attention as a novel material with well-defined porosities, high surface areas, and chemical stability [30,31]. Owing to the excellent characteristics, various biosensors based on MOFs have been developed for detecting heavy metals [29,32], antibiotics [33,34], phosphoprotein [35], and so on. However, MOFs also have some disadvantages, like the non-conducting of MOFs, as well as the unstable bond of inorganic metal ions and organic ligands. Therefore, the research of MOFs gradually transfer from the MOFs to MOF-based derivatives. Derivatives of MOFs, including metal/metal oxide nanoparticles [36], carbon−metal/metal oxide hybrids [37], and porous carbon [38] can be obtained through the solvothermal method and high-temperature calcination method [39]. Metal/metal oxide and carbon nanoparticles have been revealed excellent electrocatalytic activity for the oxygen reduction reaction. However, few studies about MOFs derivatives have been done on the electrochemical sensors.
In this work, a facile strategy on preparing carbon material (CZIFs) derived from the core–shell zeolite imidazole frameworks (ZIF8@ZIF67) was reported. The CZIFs are carbon-based hybrid double-shelled structure of outer shells of Co-graphitic carbon and inner shells of microporous carbon. The CZIFs and aptamer can generate high immobilization force for the aptamer strands. Owing to this, a new aptasensor was built to detect Pb2+ of river water and polluted soil.

2. Materials and Methods

2.1. Materials and Apparatus

The lead-aptamer (Apt) 5′-CAACGGTGGGTGTGGTTGG-3′ was synthesized by Sangon Biotechnology (Shanghai, China). PbCl2 was bought from Merck Co., Inc. (Darmstadt, Germany) and the other reagents were purchased from Shanghai Sinopharm Chemical ReagentCo., Ltd. (Shanghai, China) without further purification.
Material characterization scanning electron microscope (SEM), transmission electron microscopy (TEM), and XPS analysis were conducted on Hitachi S4800 microscopy, TF20 (FEI) microscopy, and Thermo ESCALAB 250XI, respectively. An electrochemical workstation (CHI1030A, Shanghai Chenhua Instruments, Shanghai, China) was used to measure the current. The screen printed carbon electrode (SPCE) was used in this aptasensor. The SPCE is a three electrode system including a carbon working electrode, an Ag/AgCl reference electrode, and a carbon auxiliary electrode. The full automatic microwave (Auto Digiblock S60 UP, Lab Tech) was employed to digest the polluted soil and the graphite furnace atomic absorption spectrometry (GFAAS, AA900, PE, Winter Street Waltham, MA, USA) was applied to measure the content of Pb2+.

2.2. The synthesis of CZIFs Particles

ZIF-8 particle was assembled according to previous reports [39]. Furthermore, the detailed preparation processes of the CZIFs material is shown in the Supplementary Information (Figures S1 and S2).

2.3. Preparation of Probe

One milligram of CZIFs was dissolved into 1 mL of 1% chitosan solution and ultrasonicated for 5 min. Then, the homogeneous solution was dropped onto the surface of the working electrode and dried at 25 °C. Thereafter, 5.0 μM Pb2+ aptamer was evenly introduced on the surface of the working electrode. After it dried, the modified electrodes was rinsed carefully with ultrapure water. Then, the immobilized electrodes were incubated with Pb2+ (2 μg L−1) for 30 min. After this, the electrodes were rinsed with ultrapure water again. Finally, the cyclic voltammetry (CV) was measured in the HAc-NaAc-thionine buffer of the pH 5.5, scanning from –0.65 V to 0.25 V at a scan rate of 50 mV/s. Differential Pulse Voltammetry (DPV) was performed in 0.1 M HAc-NaAc-thionine scanning in the range of –0.5 to –0.1 V at 50 mV amplitude and 0.5 s pulse width. The preparation process for electrochemical aptasensor and the detecting Pb2+ are shown in Scheme 1.

2.4. Application in aqueous samples

The river water was collected from Huangpu River, and some cations were added to detect the recovery of the aptasensor. The collection and pretreatment of polluted soil was the same with another paper [40].

3. Results and Discussion

3.1. Characterization of Porous Carbon Materials

The uniform polyhedral nanocrystals of ZIF-8 and ZIF-67 were synthesized. As shown in Figure S1, the volume of ZIF-8 nanocrystals are smaller than ZIF-67 so that the ZIF-8 can be used as the cores for epitaxial growth of ZIF-67. The SEM image of ZIF-8@ZIF-67 presented that the sample displayed entirely uniform rhombic with a smooth surface and the average edge length of the particles is about 420 nm (Figure 1).
As shown in Figure 2, the core–shell-structure particles of ZIF-8@ZIF-67 become Metal–Organic-Framework-Derived Hybrid Carbon (CZIFs) by thermal annealing. The SEM image indicated that the CZIFs samples retained crystallite shapes similar to the ZIF-8@ZIF-67 precursors, but the particles have a rough surface anchored with amounts of short carbon nanotubes (Figure 2a,b). Moreover, the high-temperature carbonization process also led to a bit of shrinkage and disintegration. The average edge length is about 420 nm and the outer shells consist of relatively loose carbon with an average thickness of around 70 nm (Figure 2b,c). The formation of short carbon nanotubes is due to the catalytic effect of Co nanoparticles on the particle surface [41]. The image of TEM presented the formation of graphitic carbon structure was decorated with Co nanoparticles and was enclosed in several graphitic carbons with a length of 5 to 15 nm (Figure 2c). The Co nanoparticles in graphitic carbon sheath are not only very stable but can also enhance the stability [42,43]. Since ZIF-67 is less stable than ZIF-8, it would first resolve to form rigid shells of CoO and carbon composites which can generate an outward adhesive force at the interface when the temperature is above 500 °C. The adhesive force is important for keeping the structure of ZIFs because it can prevent the collapse caused by the ZIF-8 induced polycondensation. Then with temperature rise, the outer shells of CoO was reduced to metallic Co with the surrounding carbonaceous become to graphitic carbon [38]. The XPS of CZIFs revealed that the main elementals of the nanoparticles are C, N, O, Co, Zn. Moreover, compared to ZIF-8@ZIF-6, the content of Co and Zn was reduced. One reason is that the unstable Co was washed by diluted sulphuric acid [39]. Another reason is, with the temperature rise, the ZIF-8 decomposed to ZnO and carbon composite and subsequent reduction to Zn and carbon composite when the temperature is above 600 °C, following the Zn to vaporize [42].

3.2. Electrochemical Characterization of the Aptasensor

In this aptasensor, the electrochemical behaviors in HAc-NaAc-thionine buffer solution after each modification step were measured by CV (Figure 3). At the bare working electrode, the redox peak of thionine revealed a minimum peak current (Figure 3a), and the peak current increased significantly after the CZIFs were introduced (curve b). The peak current decreased when the working electrode was modified by aptamer (curve c). This phenomenon is that the aptamers are bent, folded, long, and non-conductive oligonucleotide, which can block the electron transfer. In the presence of Pb2+, the peak current increased continuously (curve d), suggesting that the aptamer had become the G-quadruplex when Pb2+ was introduced. As a substitute, the lead(II)-induced allosteric G-quadruplex oligonucleotide is now being used as a functional DNA molecule for Pb2+ sensing [44,45]. It was confirmed that the presence of Pb2+ could generate the folding of Aptmer immobilized onto the ERGO/GCE electrode to a G-quadruplex structure [46].
Some reasons can be used to explain the phenomenon. As is known, the aptamer is a bent, folded, long, and non-conductive oligonucleotide. Thionine binds via external stacking to the single strand DNA (ssDNA), and DNA quadruplex only by Coulombic interaction. The electrical conductivity is not affected. In contrast, it intercalates binding to double-strand DNA (dsDNA) base pairs and the electrical conductivity can weaken [47]. In addition, the binding affinities of thionine were remarkably higher with dsDNA compared to ssDNA [48]. Aptamers of Pb2+ displayed the double stranded ends, which can bind more thionine, so that the current was decreased with the oligonucleotide introduced. While in the presence of Pb2+, the structure of aptamer becomes to G-quadruplex binding thionine by Coulombic interaction which can enhance the electric conductivity. Furthermore, the current was increased with the increase of Pb2+. On another hand, some research showed that carbon nanotube could be combined with thionine by π–π reaction which improved the conductive ability [49]. With the aptamer becoming G-quadruplex, a part of nucleic acids left the working electrode so that thionine can be combined with electrode directly [50,51] and accelerated electron transfer under the oxidation and reduction process.

3.3. Optimization of Experimental Conditions

In order to balance the factors of the proposed aptasensor, some experimental parameters affecting the performance, including a concentration of CZIFs and aptamer, a reaction time of aptasensor with Pb2+, and pH of buffer solution, were optimized (Figure S4). The results of the optimal experimental conditions are below: concentration of CZIFs: 0.5 mg mL−1, (b) concentration of aptamer: 0.5 μM, (c) pH of buffer solution: 5.5, (d) reaction time of Pb2+: 40 min.

3.4. Performance of the Aptasensor

The DPV was carried out to describe the performance of this aptasensor. Under the most appropriate conditions, difference concentrations of Pb2+ were determined. As shown in Figure 4a, the current values increased with the increasing concentration of Pb2+. The ΔI increased linearly in the range of 0.1–20 μg L−1 of Pb2+ and the linear regression equation was ΔI = 0.46C + 2.59 (R2 = 0.98, Figure 4b) with the limit of detection (LOD) of 0.096 μg L−1 (S/N = 3), where C represents the Pb2+ concentration and ΔI is the peak current response (ΔI = Ip’–Ip) before (Ip) and after (Ip’) the target of Pb2+ treatment.
The different methods of range and LOD were compared, and the results are shown in Table 1. Compared with other methods, the LOD of this aptasensor was lower and there are no significant differences with the other methods. However, the detection range of this work still needs to improve.

3.5. Selectivity and Reproducibility of the Aptasensor

Selectivity is a significant feature for an excellent aptasensor. To evaluate the selectivity of the aptasensor, some heavy metal ions which may interfere with the detection of Pb2+ were examined respectively. The concentration of Pb2+ was 2 μg L−1, the concentrations of Cu2+, Ni2+, Zn2+ were 20 μg L−1, and the concentrations of Ag+, Mn2+, Ca2+, Fe3+, K+, NH4+ were 10 μg L−1. As illustrated in Figure 5a, the other metal cations ΔI’ represented the difference between the ΔI of blank and Pb2+ and only changed slightly compared to the blank treatment. As expected, the ΔI’ of Pb2+ was higher than other cations, but the high concentrations of K+, NH4+, and Ag+ still should be studied in the future.
Moreover, the reproducibility was also significant for the aptasensor. Five equal electrodes were used to invest the reproducibility under the same experimental conditions. As shown in Figure 5b, standard deviation of five independent measurements of Pb2+ was 4.04%, indicating that the proposed aptasensor shows good reproducibility.

3.6. Assay of Pb2+ in River Water and Soil Sample

The application of the proposed assay in actual samples was explored by river water with Pb2+ and some other metal cations. The results are shown in Table 2, the mean recoveries were in the range of 96.2~101.5%, and the relative standard deviations (RSD) were below 5.0%.
The digested solutions of polluted soils were detected by proposed methods and the presently available method of GFAAS to investigate the feasibility and accuracy of the sensing platform in real samples. As shown in Figure 6, the contents of Pb2+ measured by proposed methods and GFAAS were 12.4 mg kg−1 and 12.0 mg kg−1 in dry soil, respectively. It shows the applicability of the aptasensor as a quantitative method in actual samples.

4. Conclusions

In conclusion, an electrochemical aptasensor for rapid and specificity detection of Pb2+ was successfully developed. The aptasensor was fabricated after the aptamer was immobilized on SPCE with CZIFs. In the method, CZIFs was used to provide signal transduction and amplification, and the thionine worked as a signal factor. Under optimal conditions, it was demonstrated that the aptasensor could be used to detect Pb2+ over a range of concentrations (0.1~10 μg L−1) with the LOD of 0.096 μg L−1. The method also showed high sensitivity and good selectivity. The feasibility of determining spiked water and polluted soil samples was investigated. This method shows practicality for Pb2+ detection in practice. There are some limitations in that K+, NH4+, and Ag+ of high concentration may affect the performance of the way, and the work on how to eliminate the effects would be carried out in the future.

Supplementary Materials

The following are available online at https://www.mdpi.com/2079-6374/11/1/1/s1, Figure S1: SEM images of (a) ZIF-8 and (b) ZIF-67; Figure S2: Optimization of the experimental parameters: (a) concentration of porous carbon, (b) concentration of aptamer, (c) pH of buffer solution, (d) different incubation time on aptasensor with Pb2+.

Author Contributions

Conceptualization, J.D.; methodology, J.D. and Y.L. (Yitong Lu); software, D.Z. (Dongwei Zhang); funding acquisition, P.Z. and D.Z. (Dan Zhang); formal analysis, X.Z. and Y.L. (Yang Liu); writing—original draft preparation, J.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Shanghai Agriculture Applied Technology Development Program, China (Grant No.T20180413), Young Elite Scientists Sponsorship Program by CAST [grant numbers 2017QNRC001], Sponsored by “Chenguang Program” supported by Shanghai Education Development Foundation and Shanghai Municipal Education Commission [grant numbers 17CG07].

Conflicts of Interest

The authors declare that they have no conflict of interest.

References

  1. Jaishankar, M.; Tseten, T.; Anbalagan, N.; Mathew, B.B.; Beeregowda, K.N. Toxicity, mechanism and health effects of some heavy metals. Interdiscip. Toxicol. 2014, 7, 60–72. [Google Scholar] [CrossRef] [Green Version]
  2. Rai, P.K.; Lee, S.S.; Zhang, M.; Tsang, Y.F.; Kim, K.-H. Heavy metals in food crops: Health risks, fate, mechanisms, and management. Environ. Int. 2019, 125, 365–385. [Google Scholar] [CrossRef]
  3. Rai, P.K.; Lee, J.; Kailasa, S.K.; Kwon, E.E.; Tsang, Y.F.; Ok, Y.S.; Kim, K.-H. A critical review of ferrate(VI)-based remediation of soil and groundwater. Environ. Res. 2018, 160, 420–448. [Google Scholar] [CrossRef]
  4. Yuen, L.H.; Franzini, R.M.; Wang, S.; Crisalli, P.; Singh, V.; Jiang, W.; Kool, E.T. Pattern-Based Detection of Toxic Metals in Surface Water with DNA Polyfluorophores. Angew. Chem. Edit. 2014, 53, 5361–5365. [Google Scholar] [CrossRef] [Green Version]
  5. Budnik, L.T.; Casteleyn, L. Mercury pollution in modern times and its socio-medical consequences. Sci. Total Environ. 2019, 654, 720–734. [Google Scholar] [CrossRef] [PubMed]
  6. Zhou, H.; Yang, W.; Zhou, X.; Liu, L.; Gu, J.; Wang, W.; Zou, J.; Tian, T.; Peng, P.; Liao, B. Accumulation of Heavy Metals in Vegetable Species Planted in Contaminated Soils and the Health Risk Assessment. Int. J. Environ. Res. Public Health 2016, 13, 289. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Al-Saleh, I.; Al-Rouqi, R.; Elkhatib, R.; Abduljabbar, M.; Al-Rajudi, T. Risk assessment of environmental exposure to heavy metals in mothers and their respective infants. Int. J. Hyg. Environ. Health 2017, 220, 1252–1278. [Google Scholar] [CrossRef]
  8. El-Kady, A.A.; Abdel-Wahhab, M.A. Occurrence of trace metals in foodstuffs and their health impact. Trends Food Sci. Technol. 2018, 75, 36–45. [Google Scholar] [CrossRef]
  9. Tombelli, S.; Minunni, M.; Mascini, A. Analytical applications of aptamers. Biosens. Bioelectron. 2005, 20, 2424–2434. [Google Scholar] [CrossRef]
  10. Hamula, C.L.A.; Zhang, H.; Li, F.; Wang, Z.; Le, X.C.; Li, X.-F. Selection and analytical applications of aptamers binding microbial pathogens. Trac. Trends Anal. Chem. 2011, 30, 1587–1597. [Google Scholar] [CrossRef]
  11. Qian, Z.S.; Shan, X.Y.; Chai, L.J.; Chen, J.R.; Peng, H. A fluorescent nanosensor based on graphene quantum dots-aptamer probe and graphene oxide platform for detection of lead (II) ion. Biosens. Bioelectron. 2015, 68, 225–231. [Google Scholar] [CrossRef] [PubMed]
  12. Tang, Y.; Hu, H.; Zhang, M.G.; Song, J.; Nie, L.; Wang, S.; Niu, G.; Huang, P.; Lu, G.; Chen, X. An aptamer-targeting photoresponsive drug delivery system using “off-on” graphene oxide wrapped mesoporous silica nanoparticles. Nanoscale 2015, 7, 6304–6310. [Google Scholar] [CrossRef] [PubMed]
  13. Yang, D.; Liu, X.; Zhou, Y.; Luo, L.; Zhang, J.; Huang, A.; Mao, Q.; Chen, X.; Tang, L. Aptamer-based biosensors for detection of lead(II) ion: A review. Anal. Methods UK 2017, 9, 1976–1990. [Google Scholar] [CrossRef]
  14. Li, M.; Zhou, X.; Ding, W.; Guo, S.; Wu, N. Fluorescent aptamer-functionalized graphene oxide biosensor for label-free detection of mercury(II). Biosens. Bioelectron. 2013, 41, 889–893. [Google Scholar] [CrossRef]
  15. Zhan, S.; Xu, H.; Zhang, D.; Xia, B.; Zhan, X.; Wang, L.; Lv, J.; Zhou, P. Fluorescent detection of Hg2+ and Pb2+ using GeneFinder™ and an integrated functional nucleic acid. Biosens. Bioelectron. 2015, 72 (Suppl. C), 95–99. [Google Scholar] [CrossRef]
  16. Qi, Y.; Xiu, F.-R.; Yu, G.; Huang, L.; Li, B. Simple and rapid chemiluminescence aptasensor for Hg2+ in contaminated samples: A new signal amplification mechanism. Biosens. Bioelectron. 2017, 87, 439–446. [Google Scholar] [CrossRef]
  17. Cao, J.T.; Yang, J.J.; Zhao, L.Z.; Wang, Y.L.; Wang, H.; Liu, Y.M.; Ma, S.H. Graphene oxide@gold nanorods-based multiple-assisted electrochemiluminescence signal amplification strategy for sensitive detection of prostate specific antigen. Biosens. Bioelectron. 2018, 99 (Suppl. C), 92–98. [Google Scholar] [CrossRef]
  18. Taghdisi, S.M.; Danesh, N.M.; Ramezani, M.; Alibolandi, M.; Abnous, K. Voltammetric determination of lead(II) by using exonuclease III and gold nanoparticles, and by exploiting the conformational change of the complementary strand of an aptamer. Microchim. Acta 2017, 184, 2783–2790. [Google Scholar] [CrossRef]
  19. Yuan, M.; Song, Z.; Fei, J.; Wang, X.; Xu, F.; Cao, H.; Yu, J. Aptasensor for lead(II) based on the use of a quartz crystal microbalance modified with gold nanoparticles. Microchim. Acta 2017, 184, 1397–1403. [Google Scholar] [CrossRef]
  20. Wang, N.; Lin, M.; Dai, H.; Ma, H. Functionalized gold nanoparticles/reduced graphene oxide nanocomposites for ultrasensitive electrochemical sensing of mercury ions based on thymine-mercury-thymine structure. Biosens. Bioelectron. 2016, 79, 320–326. [Google Scholar] [CrossRef]
  21. Kulikova, T.; Gorbatchuk, V.; Stoikov, I.; Rogov, A.; Evtugyn, G.; Hianik, T. Impedimetric Determination of Kanamycin in Milk with Aptasensor Based on Carbon Black-Oligolactide Composite. Sensors 2020, 20, 4738. [Google Scholar] [CrossRef] [PubMed]
  22. Rong, Q.; Han, H.; Feng, F.; Ma, Z. Network nanostructured polypyrrole hydrogel/Au composites as enhanced electrochemical biosensing platform. Sci. Rep. 2015, 5, srep11440. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Tianyu, H.; Xu, Y.; Weidan, N.; Xingguang, S. Aptamer-based aggregation assay for mercury(II) using gold nanoparticles and fluorescent CdTe quantum dots. Microchim. Acta 2016, 183, 2131–2137. [Google Scholar] [CrossRef]
  24. Wang, C.; Qian, J.; Wang, K.; Yang, X.; Liu, Q.; Hao, N.; Wang, C.; Dong, X.; Huang, X. Colorimetric aptasensing of ochratoxin A using Au@Fe3O4 nanoparticles as signal indicator and magnetic separator. Biosens. Bioelectron. 2016, 77 (Suppl. C), 1183–1191. [Google Scholar] [CrossRef] [PubMed]
  25. Yang, S.; Song, X.; Zhang, P.; Gao, L. Heating-Rate-Induced Porous α-Fe2O3 with Controllable Pore Size and Crystallinity Grown on Graphene for Supercapacitors. ACS Appl. Mater. Inter. 2015, 7, 75–79. [Google Scholar] [CrossRef]
  26. Liu, J.; Xu, X.; Hu, R.; Yang, L.; Zhu, M. Uniform hierarchical Fe3O4@ polypyrrole nanocages for superior lithium ion battery anodes. Adv. Energy Mater. 2016, 6, 1600256. [Google Scholar] [CrossRef]
  27. Guo, X.; Wen, F.; Qiao, Q.; Zheng, N.; Saive, M.; Fauconnier, M.-L.; Wang, J. A Novel Graphene Oxide-Based Aptasensor for Amplified Fluorescent Detection of Aflatoxin M-1 in Milk Powder. Sensors 2019, 19, 3840. [Google Scholar] [CrossRef] [Green Version]
  28. Chen, M.; Gan, N.; Zhou, Y.; Li, T.; Xu, Q.; Cao, Y.; Chen, Y. A novel aptamer- metal ions- nanoscale MOF based electrochemical biocodes for multiple antibiotics detection and signal amplification. Sens. Actuators B Chem. 2017, 242, 1201–1209. [Google Scholar] [CrossRef]
  29. Zhang, Z.; Ji, H.; Song, Y.; Zhang, S.; Wang, M.; Jia, C.; Tian, J.-Y.; He, L.; Zhang, X.; Liu, C.-S. Fe (III)-based metal–organic framework-derived core–shell nanostructure: Sensitive electrochemical platform for high trace determination of heavy metal ions. Biosens. Bioelectron. 2017, 94, 358–364. [Google Scholar] [CrossRef]
  30. Mossine, A.V.; Mayhan, C.M.; Fowler, D.A.; Teat, S.J.; Deakyne, C.A.; Atwood, J.L. Zinc-seamed pyrogallol [4] arene dimers as structural components in a two-dimensional MOF. Chem. Sci. 2014, 5, 2297–2303. [Google Scholar] [CrossRef]
  31. Chen, W.H.; Yu, X.; Liao, W.C.; Sohn, Y.S.; Cecconello, A.; Kozell, A.; Nechushtai, R.; Willner, I. ATP-Responsive Aptamer-Based Metal–Organic Framework Nanoparticles (NMOFs) for the Controlled Release of Loads and Drugs. Adv. Funct. Mater. 2017, 27, 1702102. [Google Scholar] [CrossRef]
  32. Cui, L.; Wu, J.; Li, J.; Ju, H. Electrochemical Sensor for Lead Cation Sensitized with a DNA Functionalized Porphyrinic Metal–Organic Framework. Anal. Chem. 2015, 87, 10635–10641. [Google Scholar] [CrossRef] [PubMed]
  33. Zhu, H.-B.; Sun, Z.-Y. Aqueous detection of antibiotics with a Cd(II)-based metal-organic framework constructed by a tetra(1,2,4-triazole)-functionalized-bis (triphenylamine) ligand. Inorg. Chem. Commun. 2018, 96, 202–205. [Google Scholar] [CrossRef]
  34. Xing, P.; Wu, D.; Chen, J.; Song, J.; Mao, C.; Gao, Y.; Niu, H. A Cd-MOF as a fluorescent probe for highly selective, sensitive and stable detection of antibiotics in water. Analyst 2019, 144, 2656–2661. [Google Scholar] [CrossRef] [PubMed]
  35. Zhang, G.Y.; Zhuang, Y.H.; Shan, D.; Su, G.F.; Cornier, S.; Zhang, X.J. Zirconium-Based Porphyrinic Metal-Organic Framework (PCN-222): Enhanced Photoelectrochemical Response and Its Application for Label-Free Phosphoprotein Detection. Anal. Chem. 2016, 88, 11207–11212. [Google Scholar] [CrossRef] [PubMed]
  36. Yu, D.; Wu, B.; Ge, L.; Wu, L.; Wang, H.; Xu, T. Decorating nanoporous ZIF-67-derived NiCo2O4 shells on a Co3O4 nanowire array core for battery-type electrodes with enhanced energy storage performance. J. Mater. Chem. A 2016, 4, 10878–10884. [Google Scholar] [CrossRef]
  37. Zhang, J.; Yan, X.; Hu, X.; Feng, R.; Zhou, M. Direct carbonization of Zn/Co zeolitic imidazolate frameworks for efficient adsorption of Rhodamine B. Chem. Eng. J. 2018, 347, 640–647. [Google Scholar] [CrossRef]
  38. Liu, S.; Wang, Z.; Zhou, S.; Yu, F.; Yu, M.; Chiang, C.Y.; Zhou, W.; Zhao, J.; Qiu, J. Metal–Organic-Framework-Derived Hybrid Carbon Nanocages as a Bifunctional Electrocatalyst for Oxygen Reduction and Evolution. Adv. Mater. 2017, 29, 1700874. [Google Scholar] [CrossRef] [Green Version]
  39. Sun, C.; Yang, J.; Rui, X.; Zhang, W.; Yan, Q.; Chen, P.; Huo, F.; Huang, W.; Dong, X. MOF-directed templating synthesis of a porous multicomponent dodecahedron with hollow interiors for enhanced lithium-ion battery anodes. J. Mater. Chem. A 2015, 3, 8483–8488. [Google Scholar] [CrossRef]
  40. Ding, J.; Liu, Y.; Zhang, D.; Yu, M.; Zhan, X.; Zhang, D.; Zhou, P. An electrochemical aptasensor based on gold@ polypyrrole composites for detection of lead ions. Microchim. Acta 2018, 185, 7. [Google Scholar] [CrossRef]
  41. Liu, Y.Y.; Jiang, H.L.; Zhu, Y.H.; Yang, X.L.; Li, C.Z. Transition metals (Fe, Co, and Ni) encapsulated in nitrogen-doped carbon nanotubes as bi-functional catalysts for oxygen electrode reactions. J. Mater. Chem. A 2016, 4, 1694–1701. [Google Scholar] [CrossRef]
  42. Wu, H.B.; Wei, S.Y.; Zhang, L.; Xu, R.; Hng, H.H.; Lou, X.W. Embedding Sulfur in MOF-Derived Microporous Carbon Polyhedrons for Lithium-Sulfur Batteries. Chem. Eng. J. 2013, 19, 10804–10808. [Google Scholar] [CrossRef] [PubMed]
  43. Xia, B.Y.; Yan, Y.; Li, N.; Wu, H.B.; Lou, X.W.; Wang, X. A metal-organic framework-derived bifunctional oxygen electrocatalyst. Nat. Energy 2016, 1, 15006. [Google Scholar] [CrossRef]
  44. Zhan, S.; Wu, Y.; Liu, L.; Xing, H.; He, L.; Zhan, X.; Luo, Y.; Zhou, P. A simple fluorescent assay for lead(II) detection based on lead(II)-stabilized G-quadruplex formation. RSC Adv. 2013, 3, 16962–16966. [Google Scholar] [CrossRef]
  45. Taghdisi, S.M.; Danesh, N.M.; Lavaee, P.; Ramezani, M.; Abnous, K. An electrochemical aptasensor based on gold nanoparticles, thionine and hairpin structure of complementary strand of aptamer for ultrasensitive detection of lead. Sens. Actuators B Chem. 2016, 234, 462–469. [Google Scholar] [CrossRef]
  46. Yu, S.H.; Lee, C.-S.; Kim, T.H. Electrochemical Detection of Ultratrace Lead Ion through Attaching and Detaching DNA Aptamer from Electrochemically Reduced Graphene Oxide Electrode. Nanomaterials 2019, 9, 817. [Google Scholar] [CrossRef] [Green Version]
  47. Hecht, C.; Friedrich, J.; Chang, T.C. Interactions of thionin with DNA strands: Intercalation versus external stacking. J. Phys. Chem. B 2004, 108, 10241–10244. [Google Scholar] [CrossRef]
  48. Paul, P.; Kumar, G.S. Thionine Interaction to DNA: Comparative Spectroscopic Studies on Double Stranded Versus Single Stranded DNA. J. Fluoresc. 2012, 22, 71–80. [Google Scholar] [CrossRef]
  49. Ma, K.; Wang, H.; Li, H.; Wang, S.; Li, X.; Xu, B.; Tian, W. A label-free aptasensor for turn-on fluorescent detection of ATP based on AIE-active probe and water-soluble carbon nanotubes. Sens. Actuators B Chem. 2016, 230, 556–558. [Google Scholar] [CrossRef]
  50. Gao, F.; Gao, C.; He, S.; Wang, Q.; Wu, A. Label-free electrochemical lead (II) aptasensor using thionine as the signaling molecule and graphene as signal-enhancing platform. Biosens. Bioelectron. 2016, 81, 15–22. [Google Scholar] [CrossRef]
  51. Liao, Q.G.; Wei, B.H.; Luo, L.G. Aptamer based fluorometric determination of kanamycin using double-stranded DNA and carbon nanotubes. Microchim. Acta 2017, 184, 627–632. [Google Scholar] [CrossRef]
  52. Hao, T.; Zhang, C.; Lin, H.; Wei, W.; Yang, F.; Wu, Y.; Niu, L.; Kang, W.; Guo, Z. A One-Step Dual-Mode Aptasensor for Subnanomolar Detection of Lead Ions Based on Electrochemiluminescence and Fast Scan Voltammetry. J. Electrochem. Soc. 2020, 167, 126506. [Google Scholar] [CrossRef]
  53. Ran, G.; Wu, F.; Ni, X.; Li, X.; Li, X.; Liu, D.; Sun, J.; Xie, C.; Yao, D.; Bai, W. A novel label-free electrochemical aptasensor with one-step assembly process for rapid detection of lead (II) ions. Sens. Actuators B Chem. 2020, 320, 128326. [Google Scholar] [CrossRef]
  54. Simionca, I.-M.; Arvinte, A.; Ardeleanu, R.; Pinteala, M. Siloxane-Crown Ether Polyamide Based Electrode for Electrochemical Determination of Lead(II) in Aqueous Solution. Electroanalysis 2012, 24, 1995–2004. [Google Scholar] [CrossRef]
Scheme 1. Presentation diagram of the preparation process for the aptasensor and Pb2+ detection.
Scheme 1. Presentation diagram of the preparation process for the aptasensor and Pb2+ detection.
Biosensors 11 00001 sch001
Figure 1. (a) SEM image and (b) TEM image of ZIF-8@ZIF-67.
Figure 1. (a) SEM image and (b) TEM image of ZIF-8@ZIF-67.
Biosensors 11 00001 g001
Figure 2. (a) SEM image and (b,c) TEM images of porous carbon, (d) XPS full scan of ZIF-8@ZIF-67 and CZIFs.
Figure 2. (a) SEM image and (b,c) TEM images of porous carbon, (d) XPS full scan of ZIF-8@ZIF-67 and CZIFs.
Biosensors 11 00001 g002
Figure 3. Cyclic voltammetry of different electrodes in testing solution: (a) bare electrode, (b) CZIFs modified electrode, (c) CZIFs/aptamer modified electrode, (d) CZIFs/aptamer/Pb2+.
Figure 3. Cyclic voltammetry of different electrodes in testing solution: (a) bare electrode, (b) CZIFs modified electrode, (c) CZIFs/aptamer modified electrode, (d) CZIFs/aptamer/Pb2+.
Biosensors 11 00001 g003
Figure 4. (a) Differential Pulse Voltammetry (DPV) responses to aptasensor capturing a series of concentrations of Pb2+ from 0.1 to 20 μg L−1, (b) calibration curve of ΔI and Pb2+ concentration.
Figure 4. (a) Differential Pulse Voltammetry (DPV) responses to aptasensor capturing a series of concentrations of Pb2+ from 0.1 to 20 μg L−1, (b) calibration curve of ΔI and Pb2+ concentration.
Biosensors 11 00001 g004
Figure 5. (a) Specificity of the aptasensor for Pb2+, (b) independent detection of the Pb2+.
Figure 5. (a) Specificity of the aptasensor for Pb2+, (b) independent detection of the Pb2+.
Biosensors 11 00001 g005
Figure 6. Relative current signal of the polluted soil.
Figure 6. Relative current signal of the polluted soil.
Biosensors 11 00001 g006
Table 1. Comparison of different sensors for determination of Pb2+.
Table 1. Comparison of different sensors for determination of Pb2+.
MethodLinear Range (μg L−1) Detection Limit (μg L−1)Reference
Probe Fe3O4@Au-ssDNA&Ru-NH220.7–20,70020.7[52]
The screen-printed electrode (SPE) was immobilized by unique consecutive adenines (polyA) linker and aptamer0.1–10000.03[53]
Electrochemical aptasensor based on Au@ppy and aptamer 0.1–2.00.072[40]
PDMS-PA-DB18C6 electrodes were evaluated for anodic stripping voltammetric20–7003.5[54]
Electrochemical aptasensor based on Metal–Organic-Framework-Derived Hybrid Carbon and aptamer 0.1–100.096This work
Table 2. Recovery of Pb2+ from water samples (n = 4).
Table 2. Recovery of Pb2+ from water samples (n = 4).
Samples
(μg L−1)
Mean Found (μg L−1)Mean Recovery (%)RSD (%)
Pb2+ (0.5), Hg2+ (2), K+ (3.28), Zn2+ (15.6), Cu2+ (24)0.4896.24.26
Pb2+ (8.0), K+ (25.6), Ca2+ (50), Mn2+ (72.7), Cd2+ (17.8)8.15101.53.95
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ding, J.; Zhang, D.; Liu, Y.; Zhan, X.; Lu, Y.; Zhou, P.; Zhang, D. An Electrochemical Aptasensor for Pb2+ Detection Based on Metal–Organic-Framework-Derived Hybrid Carbon. Biosensors 2021, 11, 1. https://doi.org/10.3390/bios11010001

AMA Style

Ding J, Zhang D, Liu Y, Zhan X, Lu Y, Zhou P, Zhang D. An Electrochemical Aptasensor for Pb2+ Detection Based on Metal–Organic-Framework-Derived Hybrid Carbon. Biosensors. 2021; 11(1):1. https://doi.org/10.3390/bios11010001

Chicago/Turabian Style

Ding, Jina, Dongwei Zhang, Yang Liu, Xuejia Zhan, Yitong Lu, Pei Zhou, and Dan Zhang. 2021. "An Electrochemical Aptasensor for Pb2+ Detection Based on Metal–Organic-Framework-Derived Hybrid Carbon" Biosensors 11, no. 1: 1. https://doi.org/10.3390/bios11010001

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