Next Article in Journal
Anti-Cancer Activity of Phytochemicals Targeting Hypoxia-Inducible Factor-1 Alpha
Previous Article in Journal
Targeting Autophagy with Natural Products as a Potential Therapeutic Approach for Cancer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

NPA-Cu2+ Complex as a Fluorescent Sensing Platform for the Selective and Sensitive Detection of Glyphosate

1
Department of Chemistry, College of Arts and Sciences, Northeast Agricultural University, Harbin 150030, China
2
Department of Applied Biology and Chemical Technology, Hong Kong Polytechnic University, Hung Hom, Hong Kong, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2021, 22(18), 9816; https://doi.org/10.3390/ijms22189816
Submission received: 23 August 2021 / Revised: 4 September 2021 / Accepted: 6 September 2021 / Published: 10 September 2021
(This article belongs to the Section Biochemistry)

Abstract

:
Glyphosate is a highly effective, low-toxicity, broad-spectrum herbicide, which is extensively used in global agriculture to control weeds and vegetation. However, glyphosate has become a potential threat to human and ecosystem because of its excessive usage and its bio-concentration in soil and water. Herein, a novel turn-on fluorescent probe, N-n-butyl-4-(3-pyridin)ylmethylidenehydrazine-1,8-naphthalimide (NPA), is proposed. It efficiently detected Cu2+ within the limit of detection (LOD) of 0.21 μM and displayed a dramatic turn-off fluorescence response in CH3CN. NPA-Cu2+ complex was employed to selectively and sensitively monitor glyphosate concentrations in real samples accompanied by a fluorescence turn-on mode. A good linear relationship between NPA and Cu2+ of glyphosate was found in the range of 10–100 μM with an LOD of 1.87 μM. Glyphosate exhibited a stronger chelation with Cu2+ than NPA and the system released free NPA through competitive coordination. The proposed method demonstrates great potential in quantitatively detecting glyphosate in tap water, local water from Songhua River, soil, rice, millet, maize, soybean, mung bean, and milk with mild conditions, and is a simple procedure with obvious consequences and no need for large instruments or pretreatment.

1. Introduction

Glyphosate (N-(phosphonomethyl) glycine) is an efficient, low-toxicity, and nonselective herbicide against perennial weeds [1]. It is widely used in economic crops or trees such as orchards, rice, maize, soybeans, wheat, and tea fields [2,3,4]. However, glyphosate has become a risk to human health because of its improper treatment and its bioconcentration in soil and water [5]. Glyphosate is a potential endocrine disruptor and exhibits adverse effects on cell cycle regulation [6,7]. It is classified as a potential carcinogen and genotoxic to humans by the International Agency for Research on Cancer (IARC) [8]. The U.S. Environmental Protection Agency set a maximum concentration of 700 μg/L for glyphosate in water [9]. Guidelines for Canadian Drinking Water Quality prescribe a limit of 280 μg/L in drinking water [10]. Therefore, establishing a convenient and reliable method for the accurate and rapid detection of glyphosate is necessary and of great urgency.
Several effective analytical strategies have been established for the detection of glyphosate. Traditional methods depend on expensive large-scale equipment, such as gas chromatography (GC), GC coupled with mass spectrometry (GC-MS), and high performance liquid chromatography (HPLC) [11,12]. To overcome the limitations of expensive instruments, tedious pretreatments, and derivatization procedures, several techniques have been proposed, such as enzyme-linked immunosorbent assay (ELISA) [13], capillary electrophoresis (CE) [14], amperometry [15], colorimetric assay [16], fluorescence spectrometry, and electrochemical sensing [17,18]. These analytical techniques overcome the disadvantages of traditional methods to a certain extent; however, there are still some limitations. For example, ELISA utilizes expensive antibodies that are sensitive to temperature and pH, and electrochemical assays generally have short life cycles. Colorimetry is obvious; however, lacks good accuracy. Hence, it is critical to develop a facile, rapid, sensitive, and efficient method to detect glyphosate in environmental samples.
Recently, fluorescence detection for glyphosate has attracted much attention because of its easy operation, rapid response, high sensitivity, etc. Fluorescence sensors such as nanoclusters [19,20,21], quantum dots [22,23,24], and metal-organic frameworks have been used to detect glyphosate [25]. Currently, copper ion complexes are widely used due to their novelty, sensitivity, low toxicity, and membrane permeability [26,27,28]. A coumarin derivative and Cu2+ complex system was developed that selectively detected glyphosate with a good linear relationship of 0.02–1.50 μg/mL [29]. Inspired by the designs of copper complexes and our continuous interest in pesticide identification [30,31,32,33,34], a new probe for Cu2+, N-n-butyl-4-(3-pyridin)ylmethylidenehydrazine-1,8-naphthalimide (NPA), was designed and synthesized (Scheme 1). NPA exhibited fluorescence quenching in the presence of Cu2+. When glyphosate was added to the NPA-Cu2+ system, the fluorescence was discovered with a color change from colorless to green. NPA-Cu2+ can be used for ultra-sensitive detection of glyphosate in environmental samples with high sensitivity, fast response, and obvious consequences.

2. Results

2.1. Spectral Characteristics of Probe NPA upon Coordination with Cu2+

The solvent effects of NPA were studied in solvents of CH3CN, CH3CH2OH, DMSO, DMF, and CH3OH, and CH3CN was selected as the test solvent (Figure S4).
The responses of NPA were examined to evaluate the sensing properties with the introduction of other different metal ions. The UV–vis absorption spectrum of free NPA showed a maximum peak at 440 nm. The absorbance intensity of NPA decreased, accompanied by the yellow solution of NPA changing to colorless after the addition of Cu2+, while the UV–vis spectra of other metal ions basically did not change, which indicated that NPA could serve as a highly selective “naked-eye” probe for Cu2+ (Figure 1a). The introduction of Cu2+ into the solution of NPA also caused significant fluorescence quenching (Figure 1b). Free NPA exhibited strong fluorescence with emission at 533 nm. The fluorescence intensity decreased quickly and was quenched with Cu2+ added after 100 s, after which the system color changed from green to colorless (Figure S5).
To further explore the characteristics of NPA for Cu2+, competition and reversibility experiments were carried out. The fluorescence intensity of the NPA-Cu2+ system was slightly enhanced after Al3+ was added, and the introduction of other interference metal ions did not affect the detection of Cu2+ by NPA (Figure S6). Alternating additions of Cu2+ and EDTA to the system caused variation in the fluorescence intensity (Figure S7). Although a certain degree of decrease in the fluorescence intensity was found after several cycles, NPA showed good reversibility. The results indicate that NPA could serve as a specific, rapid, and reversible sensor for Cu2+ detection.
The fluorescence intensity of NPA was recorded at various concentrations of Cu2+ in CH3CN for quantitative analysis. The fluorescence intensity decreased gradually with increasing the Cu2+ concentration from 0 to 1.9 eq. and remained stable after 1.9 eq. Cu2+ was added (Figure 2a). A desirable linear relationship was observed between the fluorescence intensity and Cu2+ concentration, represented by y = −21.20x + 376.29 (R2 = 0.99) (Figure 2b). The binding constant (K11) was calculated to be 1.8 × 104 M−1 (Figure S8), which was calculated with validated open-source software (BindFit) [35]. The limit of detection (LOD) for Cu2+ was calculated to be 0.21 μM based on the equation of LOD = 3σ/k, where σ is the blank standard deviation and k is the slope of the fluorescence intensity ratio vs. analyte concentration plot [36]. Comparing with NPA and other Cu2+ sensors based on Schiff’s base with different sensing mechanisms (Table S1), the LOD of NPA was low enough to detect Cu2+.
To determine the stoichiometric ratio of NPA binding to Cu2+, a Job’s plot was constructed (Figure 3). The maximum fluorescence change was observed at a mole fraction of 0.65, indicating a 2:1 binding mode of NPA and Cu2+.
The IR and NMR spectra of NPA-Cu2+ were employed to elucidate the sensing mechanism. The IR spectroscopic comparison is shown in Figure 4 and Figure 5. These results further confirm that NH participated in the complex leading to a reverse photoinduced electron transfer (PET) process and resulted in fluorescence quenching (Scheme 2).
To further explore the changes in NPA and Cu2+ before and after coordination, the frontier orbital energy diagram of NPA and NPA-Cu2+ from the PBE was constructed using DFT (Figure 6). Electrons of the fluorophore in NPA transferred from its HOMO to the LUMO and then returned to its HOMO without passing through the HOMO of the receptor in NPA, causing the free probe NPA to exhibit a fluorescence when on. When Cu2+ was added, the ΔE of the system decreased, and the electrons were transferred from the 1,8-naphthalimide moiety (HOMO-1) to the electrondeficient pyridine groups (HOMO-3). These results demonstrate that the quenching response of NPA to Cu2+ ions could be considered a reverse PET process [37,38,39].

2.2. Sensing Assay for Glyphosate by NPA-Cu2+

The binding mode between glyphosate and Cu2+ has been confirmed by EXAFS spectra [40,41]. Cu2+ ions lie at the center of a Jahn-Teller distorted octahedron with the amine, carboxylate, and phosphonate groups of glyphosate chelating with Cu2+ to form two five-membered chelate rings oriented in the equatorial plane. A sensing approach for glyphosate was developed, inspired by the aforementioned mechanism via competitive coordination between the NPA-Cu2+ complex and glyphosate, as shown in Scheme 3. When glyphosate is present in the NPA-Cu2+ system, the functional groups of glyphosate, such as carboxylate, amine, and phosphonate, chelated with Cu2+ to form the more stable complex, in order to release the NPA, and the fluorescence intensity of the system can recover. Therefore, it is feasible in theory to regard the competitive coordination balance between NPA-Cu2+ and glyphosate as a new method for the detection of glyphosate residue.
The fluorescence intensity gradually increased with increasing glyphosate concentration in the NPA-Cu2+ system (Figure 7a). The fluorescence response versus the concentrations of glyphosate showed a desirable linear relationship in the concentration range of 0 to 11 eq., as described by the linear equation y = 2.40x + 34.53, with a linear coefficient of 0.99 (Figure 7b). The binding constant (K11) was calculated as K11 = 9.6 × 104 M−1, according to Host-Guest interaction BindFit (Figure S9). The LOD was calculated to be 1.87 μM (0.32 μg/mL).
To evaluate the specificity of the NPA-Cu2+ system to glyphosate, selective and interferential tests were conducted. Some typical organophosphorus pesticides were selected, such as glufosinate, trichlorfon, phosethy-Al, trichlorfon, and fosthiazate. In addition, the triketone herbicide mesotrione, the fluorine-containing pesticide oxyfluorfen, and common everyday ligands (alanine and serine) were selected as negative interferences (Scheme 4). As shown in Figure 8a, after introducing glyphosate to the NPA-Cu2+ system, the fluorescence intensity of the system increased significantly. When glufosinate, alanine, and serine were added, the fluorescence intensity of the system slightly increased, while the other pesticides did not affect the chelation (Figure 8b).
Other pesticides did not interfere with NPA-Cu2+ system, and glyphosate was quickly and specifically recognized within 300 s (Figure S10). All of these results indicate that NPA-Cu2+ detects glyphosate specifically and quickly, and is valuable in monitoring glyphosate in real time.

2.3. Applications in Real Samples

To explore the practicality of the NPA-Cu2+ system, tap water, local water from Songhua River, soil collected from the Northeast Agricultural University campus (Harbin, China), rice, millet, maize, soybean, mung bean, and milk directly purchased from local supermarkets were used for comparison. Different concentrations (30, 60, and 90 μM) of glyphosate standard solutions were added to these samples and detected through the NPA-Cu2+ system. Satisfactory fortified recoveries of 98.8–116.7% were obtained, and the relative standard deviations were all less than 2.5% (Table 1). Compared with previously reported sensing systems for glyphosate (Table 2), this strategy is more sensitive than others, and features no enzymatic mild conditions, has no need for large instruments or pretreatment, but has a simple procedure with obvious consequences [42,43,44,45]. The results indicate that the sensor system possesses significant potential for practical applications related to the environmental monitoring of glyphosate.

3. Discussion

3.1. Sensing Mechanism of Cu2+ by NPA

There was a sharp peak at around 3309 cm−1 in the IR spectra of NPA, which could be assigned to the NH of the hydrazine group. This peak was replaced by a broad peak at 3358 cm−1 in the complex (NPA-Cu2+) spectrum, indicating the involvement of NH in the complexation. To further investigate the binding site of NPA with Cu2+, 1H NMR titrations were carried out. The obvious change was that the proton signal (site i) of the NH gradually decreased with the addition of Cu2+, while the proton signal of the =CH− group in the pyridine ring at 10.23 ppm (site h) gradually shifted to the lower field due to the paramagnetic properties of Cu2+. When 3 eq. Cu2+ was added, the proton signal of the NH almost disappeared and the proton signal of the =CH− group in the pyridine ring moved to 10.85 ppm. These further confirm that NH participated in the complex leading to a reverse photoinduced electron transfer (PET) process and resulted in fluorescence quenching [46,47,48].

3.2. Sensing Mechanism of Glyphosate by NPA-Cu2+

The experimental result that the NPA-Cu2+ complex system can specifically recognize glyphosate is attributed to the relative spatial proximity of the phosphonate amino and carboxyl groups of glyphosate to each other, facilitating the formation of chelates with Cu2+. Although glufosinate and glyphosate are similar in structure, the amino and phosphonate groups are separated by three methylenes in glufosinate, which make it difficult to form a steric chelation with Cu2+. Alanine and serine are structurally similar to glufosinate, and cannot form a stable steric chelation with Cu2+. However, other similar pesticides lacking coordination groups failed to form stable coordinations with Cu2+, and could not form chelations in space.

4. Materials and Methods

4.1. Materials and Physical Instruments

All analytical reagent-grade chemicals and solvents employed for the experiment, which were purchased from commercial providers, were used without further purification. Glyphosate, glufosinate, trichlorfon, phosethy−Al, fosthiazate, mesotrione, and oxyfluorfen were purchased from Altai Biological Technology Co., Ltd. (Hebei, China).
FT−IR spectra were measured on a Bruker ALPHA−T spectrometer, which used KBr pellets with a range of 4000–600 cm−1 (Bruker Corp., Billerica, MA, USA). The 1H and 13C NMR spectra were recorded on a Bruker AV400 NMR spectrometer by using DMSO-d6 as the solvent, and the chemical shifts are reported in ppm (Bruker Corp., Billerica, MA, USA). The HRMS was obtained on an FTMS Ultral Apex MS spectrometer (Bruker Corp., Billerica, MA, USA). The absorption spectra were gained on a Shimadzu UV-2700 UV–vis spectrometer at 25 °C (Shimadzu Corp., Kyoto, Japan). Fluorescence spectra were obtained on a PerkinElmer LS55 fluorescence spectrometer (PerkinElmer Corp., Waltham, MA, USA) with a xenon lamp and quartz carrier.

4.2. Synthesis of the Probe NPA

N-n-butyl-4-bromo-1,8-naphthalimide (1) and N-n-butyl-4-hydrazine hydrate-1,8 naphthalic anhydride (2) were synthesized according to previous research of our group [49,50,51]. Compound 2 (283.0 mg, 1 mmol) and β-nicotinaldehyde (40%, w/w, 3.0 mL) were added to EtOH (20 mL). The mixture was refluxed for 2 h and then cooled to room temperature. The precipitate was filtered and washed with EtOH to obtain an orange solid with a yield of 66%. IR (KBr, ν, cm−1) 3309(N–H), 2957, 2858(C–H), 1635(C=O), 1126(C–N). 1H NMR (400 MHz, DMSO-d6) δ 11.48 (s, 1H), 10.23 (s, 1H), 8.82 (d, J = 5.9 Hz, 1H), 8.49 (d, J = 7.3 Hz, 1H), 8.39 (d, J = 8.5 Hz, 1H), 7.85 (d, J = 7.8 Hz, 1H), 7.82–7.77 (m, 1H), 7.66 (d, J = 8.5 Hz, 1H), 7.26 (t, J = 7.7 Hz, 1H), 6.96–6.90 (m, 2H), 4.06–4.01 (m, 2H), 1.65–1.57 (m, 2H), 1.35 (m, J = 7.4 Hz, 2H), 0.93 (t, J = 7.4 Hz, 3H).13C NMR (100 MHz, DMSO-d6) δ 164.02, 163.36, 149.94, 148.27, 146.53, 140.92, 134.01, 133.69, 131.31, 131.22, 129.45, 128.61, 125.50, 124.62, 122.46, 119.13, 112.00, 107.65, 40.02, 30.23, 20.31, 14.19. HRMS (ESI): calculated for C22H20N4O2 [M + H]+ 373.1658, obtained a value of 373.1659.
The original spectra of NPA are included in the electronic Supplementary Information (Figures S1–S3).

4.3. General Procedures for Spectrophotometric Studies

The solvent effects of NPA (10−5 M) were studied in the solvents of CH3CN, CH3CH2OH, DMSO, DMF, and CH3OH, and then CH3CN was selected as the solvent to be tested. The stock solutions of 10−2 M metal ions were provided from NaCl, KCl, CuCl, AgNO3, CuCl2, MgCl2, NiCl2, ZnCl2, SnCl2, BaCl2, MnCl2, CaCl2, HgCl2, PbCl2, BaCl2, CoCl2, HgCl2, PbCl2, FeCl2, AlCl3, and CrCl3 using ultrapure water. A Cu2+ solution (1 mL, 10−2 M) was added to a 10 mL volumetric flask and diluted to 10−3 M in ultrapure water. NPA (10−5 M) solutions of 9.0, 8.0, 7.0, 6.0, 5.0, 4.0, 3.0, 2.0, and 1.0 mL were taken and transferred to 10.0 mL volumetric flasks. Cu2+ (10−3 M) solutions of 10, 20, 30, 40, 50, 60, 70, 80, and 90 µL were added to each NPA solution, so that the mole fractions of NPA became 0.1, 0.2, 0.3, 0.4, 0.5, 0.6, 0.7, 0.8, and 0.9, respectively. Each bottle was diluted with CH3CN to 10 mL [52]. Ethylene diamine tetraacetic acid (EDTA) solution (10−2 M) for a reversible experiment was obtained with ultrapure water.

4.4. Theoretical Calculation Methods

The quantum chemical calculations of the optimized structures were conducted by applying density functional theory (DFT) in Material Studio. The structure was optimized using the Perdew-Burke-Ernzerh (PBE) method of generalized gradient approximation (GGA) in the Dmol3 module for NPA and a model NPA-Cu2+ complex.

4.5. Measurement of Glyphosate Using the NPA-Cu2+ System

An NPA solution (10.0 mL, 10−4 M) and a Cu2+ aqueous solution (200 μL, 10−2 M) were added to a 100 mL volumetric flask, and then the volume was fixed with CH3CN to configure the NPA-Cu2+ solution (10−5 M). Subsequently, the final mixture was incubated for 5 min at 30 °C, and the fluorescence intensity at 533 nm was recorded with λex at 440 nm. Interference (glyphosate, glufosinate, trichlorfon, phosethy-Al, fosthiazate, mesotrione, oxyfluorfen, alanine, and serine) solutions (10−2 M) were prepared for interference and selective experiments.

4.6. Applications in Real Samples

In order to verify the practical applications, tap water, local water from Songhua River, and soil collected from the Northeast Agricultural University campus (Harbin, China), rice, millet, maize, soybean, mung bean, and milk purchased directly from the local supermarkets were selected as the real samples and were tested by the proposed sensor through the standard addition method. Water from Songhua River was filtered through a 0.22 µm membrane to remove large solids and most impurities. The rice, millet, maize, soybean, and mung bean samples were ground into powders. Soil and rice millet, maize, soybean, and mung bean samples (10 g) were dispersed in 100 mL deionized water under ultrasonication for 15 min, centrifuged at 5000 rpm for 10 min, and then filtered through a 0.22 μm microporous membrane for subsequent analyses [40,53]. Additionally, different concentrations (30, 60, and 90 μM) of glyphosate standard solutions were added to these samples and detected through the NPA-Cu2+ system.

5. Conclusions

In summary, a highly selective and sensitive naphthalimide-based derivative sensor NPA and a Cu2+ complex (NPA-Cu2+ system) for glyphosate detection were established. NPA was synthesized as an indicator of Cu2+ with turn-off fluorescence through coordination, and the LOD for Cu2+ detection was found to be 0.21 μM. The phosphonate, amino, and carboxyl groups of glyphosate are relatively close to each other, allowing a chelate to form, and thus the space cyclization to chelate with Cu2+. When glyphosate was added to the NPA-Cu2+ system, NPA was released within 300 s and the fluorescence showed a turn-on pattern with a detection limit of 1.87 μM for glyphosate. Further study revealed that glyphosate was successfully detected in water, soil, rice, millet, maize, soybean, mung bean, and milk samples by the NPA-Cu2+ system. Considering the merits of the NPA-Cu2+ system to detect glyphosate, namely it is rapid, sensitive mild, easy to operate, and does not require large-scale equipment, it is expected to become a potential method for the efficient detection of glyphosate.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/ijms22189816/s1.

Author Contributions

Y.F. conceived and designed the experiment and conveyed the manuscript information as corresponding authors. F.S. and X.-L.Y. repeated the experiment multiple times and then used it as content for in-depth research. F.S. and X.-L.Y. used Origin and Microsoft Excel to process the data. M.-L.Y. and Y.-B.W. performed theoretical calculations. P.L. coordinated all of the data. Y.-L.L. and L.Y. consulted related literature, collected background knowledge and provided theoretical support. F.S. and X.-L.Y. synthesized and purified the compound. F.S. wrote the manuscript, which was edited by Y.F. All authors have read and agreed to the published version of the manuscript.

Funding

The work was supported by the National Natural Science Foundation of China (No. 51903032), the Natural Science Foundation of Heilongjiang Province (No. LH2019B002), the Postdoctoral Science Foundation of Heilongjiang Province (No. LBH-Z19045), and the “Young Talents” Project of Northeast Agricultural University of Heilongjiang Province (No. 18QC64).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors thank Min Zhang (Northeast Normal University) for assistance with the DFT calculation and analysis.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Qiao, C.K.; Wang, C.X.; Pang, R.L.; Tian, F.J.; Han, L.J.; Guo, L.L.; Fang, J.B. Environmental behavior and influencing factors of glyphosate in peach orchard ecosystem. Ecotox. Environ. Safe 2020, 206, 11209. [Google Scholar] [CrossRef]
  2. Benbrook, C.M. Trends in glyphosate herbicide use in the United States and globally. Environ. Sci. Eur. 2016, 28, 3. [Google Scholar] [CrossRef] [Green Version]
  3. Conrad, A.; Schroter-Kermani, C.; Hoppe, H.W.; Ruther, M.; Pieper, S.; Kolossa-Gehring, M. Glyphosate in German adults-time trend (2001 to 2015) of human exposure to a widely used herbicide. Int. J. Hyg. Environ. Health 2017, 220, 8–16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Duke, S.O. The history and current status of glyphosate. Pest Manag. Sci. 2018, 74, 1027–1034. [Google Scholar] [CrossRef] [PubMed]
  5. Liu, Z.; Yang, L.; Sharma, A.S.; Chen, M.; Chen, Q. A system composed of polyethylenimine-capped upconversion nanoparticles, copper (II), hydrogen peroxide and 3,3’,5,5’-tetramethylbenzidine for colorimetric and fluorometric determination of glyphosate. Microchim. Acta 2019, 186, 835. [Google Scholar] [CrossRef]
  6. Gasnier, C.; Dumont, C.; Benachour, N.; Clair, E.; Chagnon, M.C.; Seralini, G.E. Glyphosate-based herbicides are toxic and endocrine disruptors in human cell lines. Toxicology 2009, 262, 184–191. [Google Scholar] [CrossRef] [PubMed]
  7. Wozniak, E.; Reszka, E.; Jablonska, E.; Balcerczyk, A.; Broncel, M.; Bukowska, B. Glyphosate affects methylation in the promoter regions of selected tumor suppressors as well as expression of major cell cycle and apoptosis drivers in PBMCs (in vitro study). Toxicol. In Vitro 2020, 63, 104736. [Google Scholar] [CrossRef] [PubMed]
  8. Guyton, K.Z.; Loomis, D.; Grosse, Y.; Grosse, F.E.; Benbrahim-Tallaa, L.; Guha, N.; Scoccianti, C.; Mattock, H.; Straif, K. Carcinogenicity of tetrachlorvinphos, parathion, malathion, diazinon, and glyphosate. Lancet Oncol. 2015, 16, 490–491. [Google Scholar] [CrossRef]
  9. USEPA. Edition of the Drinking Water Standards and Health Advisories; United States Environmental Protection Agency: Washington, DC, USA, 2017. [Google Scholar]
  10. Health Canada. Guidelines for canadian drinking water quality-supporting documents-glyphosate. In Health Canada Guidelines; Health Canada: Ottawa, ON, Canada, 1995. [Google Scholar]
  11. Okada, E.; Coggan, T.; Anumol, T.; Clarke, B.; Allinson, G. A simple and rapid direct injection method for the determination of glyphosate and AMPA in environmental water samples. Anal. Bioanal. Chem. 2019, 411, 715–724. [Google Scholar] [CrossRef]
  12. Islas, G.; Rodriguez, J.A.; Mendoza-Huizar, L.H.; Perez-Moreno, F.; Carrillo, E.G. Determination of glyphosate and aminomethyl phosphonic acid in soils byhplc with precolumn derivatization using 1,2-naphthoquinone-4-sulfonate. J. Liq. Chromatogr. Relat. Technol. 2014, 37, 1298–1309. [Google Scholar] [CrossRef]
  13. Mortl, M.; Nemeth, G.; Juracsek, J.; Darvas, B.; Kamp, L.; Rubio, F.; Szekacs, A. Determination of glyphosate residues in Hungarian water samples by immunoassay. Microchem. J. 2013, 107, 143–151. [Google Scholar] [CrossRef]
  14. Corbera, M.; Hidalgo, A.; Salvado, V.; Wieczorek, P. Determination of glyphosate and aminomethylphosphonic acid in natural water using the capillary electrophoresis combined with enrichment step. Anal. Chim. Acta 2005, 54, 3–7. [Google Scholar] [CrossRef]
  15. Sanchez-Bayo, F.; Hyne, R.V.; Hyne, K.L. An amperometric method for the detection of amitrole, glyphosate and its aminomethyl-phosphonic acid metabolite in environmental waters using passive samplers. Anal. Chim. Acta 2010, 675, 125–131. [Google Scholar] [CrossRef]
  16. Chen, Q.A.; Chen, H.; Li, Z.H.; Pang, J.; Lin, T.R.; Guo, L.Q.; Fu, F.F. Colorimetric sensing of glyphosate in environmental water based on peroxidase mimetic activity of MoS2 nanosheets. J. Nanosci. Nanotechnol. 2017, 17, 5730–5734. [Google Scholar] [CrossRef]
  17. Perez, A.L.; Tibaldo, G.; Sanchez, G.H.; Siano, G.G.; Marsili, N.R.; Schenone, A.V. A novel fluorimetric method for glyphosate and AMPA determination with NBD-Cl and MCR-ALS. Spectrochim. Acta A 2019, 214, 119–128. [Google Scholar] [CrossRef]
  18. Aguirre, M.D.; Aguirre, S.E.; Gomez, C.G. A Cu2+–Cu/glassy carbon system for glyphosate determination. Sensor. Actuat. B—Chem. 2019, 284, 675–683. [Google Scholar] [CrossRef]
  19. Hong, C.Y.; Ye, S.S.; Dai, C.Y.; Wu, C.Y.; Chen, L.L.; Huang, Z.Y. Sensitive and on-site detection of glyphosate based on papain-stabilized fluorescent gold nanoclusters. Anal. Bioanal. Chem. 2020, 412, 8177–8184. [Google Scholar] [CrossRef] [PubMed]
  20. Jiang, M.D.; Chen, C.; He, J.B.; Zhang, H.Y.; Xu, Z.X. Fluorescence assay for three organophosphorus pesticides in agricultural products based on Magnetic-Assisted fluorescence labeling aptamer probe. Food Chem. 2020, 307, 125534. [Google Scholar] [CrossRef]
  21. Wang, X.F.; Yang, Y.X.; Huo, D.Q.; Ji, Z.; Ma, Y.; Yang, M.; Luo, H.B.; Luo, X.G.; Hou, C.J.; Lv, J.Y. A turn-on fluorescent nanoprobe based on N-doped silicon quantum dots for rapid determination of glyphosate. Microchimica. Acta 2020, 187, 520–527. [Google Scholar] [CrossRef]
  22. Hou, J.Z.; Wang, X.F.; Wang, S.Y.; Zhang, C.; Hou, C.J.; He, Q.; Huo, D.Q. A turn-on fluorescent sensor based on carbon dots from Sophora japonica leaves for the detection of glyphosate. Anal. Methods 2020, 12, 4130–4138. [Google Scholar] [CrossRef]
  23. Bera, M.K.; Mohapatra, S. Ultrasensitive detection of glyphosate through effective photoelectron transfer between CdTe and chitosan derived carbon dot. Colloid Surface A 2020, 596, 124710. [Google Scholar] [CrossRef]
  24. Yuan, J.; Jiang, S.; Liu, J.; Yang, H.; Zhang, Y.J.; Hu, X. Fluorescent carbon dots for glyphosate determination based on fluorescence resonance energy transfer and logic gate operation. Sens. Actuat. B—Chem. 2020, 242, 545–553. [Google Scholar] [CrossRef]
  25. Chen, F.; Li, G.D.; Liu, H.; Leung, C.H.; Ma, D.L. G-quadruplex-based detection of glyphosate in complex biological systems by a time-resolved luminescent assay. Sens. Actuat. B—Chem. 2020, 320, 128393. [Google Scholar] [CrossRef]
  26. Zhao, H.; Liu, X.F.; Ma, C.B. Sensitive Fluorescence Assay for the Detection of Alkaline Phosphatase Based on a Cu2+-Thiamine System. Sensors 2021, 21, 674. [Google Scholar] [CrossRef] [PubMed]
  27. Padnya, P.; Shibaeva, K.; Arsenyev, M.; Baryshnikova, S.; Terenteva, O.; Shiabiev, I.; Khannanov, A.; Boldyrev, A.; Boldyrev, A.; Gerasimov, A.; et al. Catechol-containing schiff bases on thiacalixarene: Synthesis, copper (II) recognition, and formation of organic-inorganic copper-based materials. Molecules 2021, 26, 2334. [Google Scholar] [CrossRef] [PubMed]
  28. Hao, H.; Zhao, C.; Zheng, L.; Zhang, X.J. Evaluation of fluorescent Cu2+ probes: Instant sensing, cell permeable recognition and quantitative detection. Molecules 2021, 26, 512. [Google Scholar]
  29. Wang, X.F.; Sakinati, M.; Yang, Y.X.; Ma, Y.; Yang, M.; Luo, H.B.; Hou, C.J.; Huo, D. The construction of a CND/Cu2+ fluorescence sensing system for the ultrasensitive detection of glyphosate. Anal. Methods-UK 2020, 12, 520–527. [Google Scholar] [CrossRef]
  30. Yang, L.; Liu, Y.L.; Liu, C.G.; Ye, F.; Fu, Y. Two luminescent dye@MOFs systems as dual-emitting platforms for efficient pesticides detection. J. Hazard. Mater. 2020, 381, 120966. [Google Scholar] [CrossRef]
  31. Li, L.; Gao, S.; Yang, L.; Liu, Y.L.; Li, P.; Ye, F.; Fu, Y. Cobalt (II) complex as a fluorescent sensing platform for the selective and sensitive detection of triketone HPPD inhibitors. J. Hazard. Mater. 2021, 404, 124015. [Google Scholar] [CrossRef]
  32. Wang, Z.W.; Zhao, L.X.; Ma, P.; Ye, T.; Fu, Y.; Ye, F. Fragments Recombination, Design, Synthesis, Safener Activity and CoMFA Model of Novel Substituted Dichloroacetylphenyl Sulfonamide Derivatives. Pest Manag. Sci. 2021, 77, 1724–1738. [Google Scholar] [CrossRef]
  33. Ye, F.; Zhai, Y.; Guo, K.L.; Liu, Y.X.; Li, N.; Gao, S.; Zhao, L.X.; Fu, Y. Safeners improve maize tolerance under herbicide toxicity stress by increasing the activity of enzyme in vivo. J. Agric. Food Chem. 2019, 67, 11568–11576. [Google Scholar] [CrossRef] [PubMed]
  34. Fu, Y.; Wang, M.; Zhao, L.X.; Zhang, S.Q.; Liu, Y.X.; Guo, Y.Y.; Zhang, D.; Gao, S.; Ye, F. Design, synthesis, herbicidal activity and CoMFA of novel aryl-formyl piperidinone HPPD inhibitors. Pestic. Biochem. Physiol. 2021, 174, 104811. [Google Scholar] [CrossRef]
  35. Kampes, R.; Tepper, R.; Gorls, H.; Bellstedt, P.; Jager, M.; Schubert, U.S. Facile and Reliable Emission-Based Nanomolar Anion Sensing by Luminescent Iridium Receptors Featuring Chelating Halogen-Bonding Sites. Chem. Eur. J. 2020, 26, 14679–14687. [Google Scholar] [CrossRef] [PubMed]
  36. Ye, F.; Liang, X.M.; Xu, K.X.; Pang, X.X.; Chai, Q.; Fu, Y. A novel dithiourea-appended naphthalimide “on−off” fluorescent probe for detecting Hg2+ and Ag+ and its application in cell imaging. Talanta 2019, 200, 494–502. [Google Scholar] [CrossRef] [PubMed]
  37. Li, P.; Wang, Y. A new fluorescent sensor containing glutamic acid for Fe3+ and its resulting complex as a secondary sensor for PPi in purely aqueous solution. New. J. Chem. 2017, 41, 4234–4240. [Google Scholar] [CrossRef]
  38. Li, C.; Han, X.T.; Mao, S.S.; Aderinto, S.O.; Shi, X.K.; Shen, K.S.; Wu, H.L. A new, highly potent 1,8-naphthalimide-based fluorescence turn-off chemosensor capable of detecting Cu (II) ions in real-world water samples. Color Technol. 2018, 134, 230–239. [Google Scholar] [CrossRef]
  39. Sun, X.; Liu, Y.R.; Niu, N.; Chen, L.G. Synthesis of molecularly imprinted fluorescent probe based on biomass-derived carbon quantum dots for detection of mesotrione. Anal. Bioanal. Chem. 2019, 411, 5519–5530. [Google Scholar] [CrossRef]
  40. Wang, L.; Bi, Y.D.; Gao, J.; Li, Y.J.; Ding, H.; Ding, L. Carbon dots based turn-on fluorescent probes for the sensitive deter-mination of glyphosate in environmental water samples. RSC Adv. 2016, 89, 85820–85828. [Google Scholar] [CrossRef]
  41. Sheals, J.; Persson, P.; Hedman, B. IR and EXAFS spectroscopic studies of glyphosate protonation and copper (II) complexes of glyphosate in aqueous solution. Inorg. Chem. Front. 2001, 40, 4302–4309. [Google Scholar] [CrossRef]
  42. Noori, J.S.; Dimaki, M.; Mortensen, J.; Svendsen, W.E. Detection of glyphosate in drinking water: A fast and direct detection method without sample pretreatment. Sensors—Basel 2018, 18, 2961. [Google Scholar] [CrossRef] [Green Version]
  43. Saito, T.; Namera, A.; Tsuji, T.; Noguchi, W.; Inokuchi, S. Determination of glyphosate and glufosinate in human serum by monospin TiO extraction and liquid chromatography-tandem mass spectrometry. Acta Chromatogr. 2020, 32, 10–15. [Google Scholar] [CrossRef]
  44. Mbokana, J.G.Y.; Dedzo, G.K.; Ngameni, E. Grafting of organophilic silane in the interlayer space of acid-treated smectite: Application to the direct electrochemical. Apple Clay Sci. 2020, 188, 105513. [Google Scholar] [CrossRef]
  45. Moraes, M.P.; Goncalves, L.M.; Pereira, E.A. Determination of glyphosate and aminomethylphosphonic acid by capillary electrophoresis with indirect detection using pyridine-2,6-dicarboxylic acid or 3,5-dinitrobenzoic acid. Int. J. Environ. Anal. Chem. 2018, 98, 258–270. [Google Scholar] [CrossRef]
  46. Bao, Z.M.; Qin, C.X.; Wang, J.J.; Sun, J.; Dai, L.X.; Chen, G.Q.; Mei, F. Sensitive and selective probe for visual detection of Cu2+ based on 1, 8-naphthalimidederivative. Sens. Actuat. B—Chem. 2018, 265, 234–241. [Google Scholar] [CrossRef]
  47. Xiong, J.J.; Huang, P.C.; Zhou, X.; Wu, F.Y. A highly selective and sensitive “turn-on” fluorescent probe of Cu2+ by p-dimethylaminobenzamide-based derivative and its bioimaging in living cells. Sens. Actuat. B—Chem. 2016, 232, 673–679. [Google Scholar] [CrossRef]
  48. Zhao, J.L.; Tomiyasu, H.; Wu, C.; Cong, H.; Zeng, X.; Rahman, S. Synthesis, crystal structure and complexation behaviour study of an efficient Cu2+ ratiometric fluorescent chemosensor based on thiacalix[4]arene. Tetrahedron 2015, 71, 8521–8527. [Google Scholar] [CrossRef]
  49. Fu, Y.; Pang, X.X.; Wang, Z.Q.; Chai, Q.; Ye, F. A highly sensitive and selective fluorescent probe for determination of Cu (II) and application in live cell imaging. Spectrochim. Acta A 2019, 208, 198–205. [Google Scholar] [CrossRef]
  50. Bhorge, Y.R.; Tsai, H.T.; Huang, K.F.; Pape, A.J.; Janaki, S.N.; Yen, Y.P. A new pyrene-based Schiff-base: A selective colorimetric and fluorescent chemosensor for detection of Cu(II) and Fe(III). Spectrochim. Acta A 2014, 130, 7–12. [Google Scholar] [CrossRef]
  51. Xiong, J.J.; Huang, P.C.; Zhang, C.Y.; Wu, F.Y. Colorimetric detection of Cu2+ in aqueous solution and on the test kit by 4-aminoantipyrine derivatives. Sens. Actuat. B—Chem. 2016, 226, 30–36. [Google Scholar] [CrossRef]
  52. Wu, N.; Zhao, L.X.; Jiang, C.Y.; Li, P.; Liu, Y.L.; Fu, Y.; Ye, F. A naked-eye visible colorimetric and fluorescent chemosensor for rapid detection of fluoride anions: Implication for toxic fluorine-containing pesticides detection. J. Mol. Liq. 2020, 302, 230–239. [Google Scholar] [CrossRef]
  53. Qu, F.; Wang, F.; You, J. Dual lanthanide-probe based on coordination polymer networks for ratiometric detection of glyphosate in food samples. Food Chem. 2020, 323, 5519–5530. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Synthesis route to compound NPA.
Scheme 1. Synthesis route to compound NPA.
Ijms 22 09816 sch001
Figure 1. UV–vis (a) and fluorescence (b) spectra of NPA toward various metal cations.
Figure 1. UV–vis (a) and fluorescence (b) spectra of NPA toward various metal cations.
Ijms 22 09816 g001
Figure 2. (a) Fluorescence spectrum of NPA on the addition of different amounts of Cu2+ in a CH3CN solution; (b) the linear responses of NPA with Cu2+ concentrations.
Figure 2. (a) Fluorescence spectrum of NPA on the addition of different amounts of Cu2+ in a CH3CN solution; (b) the linear responses of NPA with Cu2+ concentrations.
Ijms 22 09816 g002
Figure 3. Job’s plot for the binding ratio between NPA and Cu2+.
Figure 3. Job’s plot for the binding ratio between NPA and Cu2+.
Ijms 22 09816 g003
Figure 4. The FT-IR spectra of NPA compound before and after the introduction of Cu2+.
Figure 4. The FT-IR spectra of NPA compound before and after the introduction of Cu2+.
Ijms 22 09816 g004
Figure 5. 1H NMR titration plots of NPA probe with Cu2+ in DMSO-d6. (a, c, d, e: naphthalimide; b, c, g, h: pyridine; f: CH-N; i: NH.
Figure 5. 1H NMR titration plots of NPA probe with Cu2+ in DMSO-d6. (a, c, d, e: naphthalimide; b, c, g, h: pyridine; f: CH-N; i: NH.
Ijms 22 09816 g005
Scheme 2. Possible sensing mechanism of NPA and Cu2+.
Scheme 2. Possible sensing mechanism of NPA and Cu2+.
Ijms 22 09816 sch002
Figure 6. Frontier orbital energy diagram and electron transfer path in NPA and NPA-Cu2+.
Figure 6. Frontier orbital energy diagram and electron transfer path in NPA and NPA-Cu2+.
Ijms 22 09816 g006
Scheme 3. Proposed sensing mechanism of NPA-Cu2+ and glyphosate.
Scheme 3. Proposed sensing mechanism of NPA-Cu2+ and glyphosate.
Ijms 22 09816 sch003
Figure 7. (a) Fluorescence spectrum of NPA-Cu2+ on the addition of different amount of glyphosate in a CH3CN solution; (b) the linear responses of NPA-Cu2+ with glyphosate concentrations.
Figure 7. (a) Fluorescence spectrum of NPA-Cu2+ on the addition of different amount of glyphosate in a CH3CN solution; (b) the linear responses of NPA-Cu2+ with glyphosate concentrations.
Ijms 22 09816 g007
Scheme 4. The structures of test interferences.
Scheme 4. The structures of test interferences.
Ijms 22 09816 sch004
Figure 8. (a) Fluorescence spectral changes in NPA-Cu2+ toward various pesticides (Gly: glyphosate); (b) fluorescence response of NPA-Cu2+ upon the addition of various pesticides.
Figure 8. (a) Fluorescence spectral changes in NPA-Cu2+ toward various pesticides (Gly: glyphosate); (b) fluorescence response of NPA-Cu2+ upon the addition of various pesticides.
Ijms 22 09816 g008
Table 1. Detection of glyphosate in water samples and soil samples using the proposed method.
Table 1. Detection of glyphosate in water samples and soil samples using the proposed method.
SamplesAdded (μM)Found (n = 3) (μM)Recovery (n = 3) (%)RSD (%)
Soil-
30
0.0 ± 0.0
34.5 ± 0.1
-
115.0 ± 1.7
-
1.9
6062.9 ± 0.2104.8 ± 0.41.2
9093.6 ± 0.1104.0 ± 0.12.0
-0.0 ± 0.0--
Songhua River3035.2 ± 0.3116.7 ± 0.22.5
6066.0 ± 0.1110.3 ± 0.11.5
9094.1 ± 0.6104.4 ± 1.10.9
-0.0 ± 0.0--
Tap Water3033.2 ± 0.2109.6 ± 1.11.6
6065.4 ± 0.4105.6 ± 0.11.4
9089.3 ± 0.598.8 ± 0.12.1
-0.0 ± 0.0--
3032.5 ± 0.1108.3 ± 0.81.5
Rice6062.8 ± 0.3104.6 ± 0.32.3
9089.5 ± 0.599.4 ± 0.51.8
-0.0 ± 0.0--
3031.9 ± 0.3106.3 ± 1.31.6
Milk6061.7 ± 0.5102.8 ± 1.51.1
9089.8 ± 0.699.7 ± 0.82.3
-0.0 ± 0.0--
3031.5 ± 0.9105.0 ± 0.30.9
Millet6061.4 ± 0.7102.5 ± 0.10.5
9090.6 ± 0.9100.7 ± 0.10.7
-0.0 ± 0.0--
3030.7 ± 0.3102.3 ± 0.30.7
Maize6060.7 ± 0.2101.2 ± 0.10.5
9090.5 ± 0.3105.5 ± 0.10.7
-0.0 ± 0.0--
3030.6 ± 0.2102.3 ± 0.30.9
Soybean6060.8 ± 0.4101.5 ± 0.51.0
9090.2 ± 0.3100.5 ± 0.61.8
-0.0 ± 0.0--
3029.0 ± 0.496.6 ± 0.10.4
Mung bean6059.2 ± 0.398.6 ± 0.60.5
9089.4 ± 0.299.3 ± 0.30.3
Table 2. Comparison of different strategies to detect glyphosate.
Table 2. Comparison of different strategies to detect glyphosate.
Sensing SystemRange (μg/mL)LOD (μg/mL)ApplicationReference
CDs–Fe3+0.1–168.75Potatoes
Water
Serum
Soil
Water
[22]
Gold Electrode50–3002.0[42]
LC-MS/MS1–2500.5[43]
Electrochemical sensor1.7–16.90.98[44]
Capillary electrophoretic methodologies1.0–8.00.5[45]
NPA-Cu2+0–18.60.32Water
Soil
This work
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Sun, F.; Ye, X.-L.; Wang, Y.-B.; Yue, M.-L.; Li, P.; Yang, L.; Liu, Y.-L.; Fu, Y. NPA-Cu2+ Complex as a Fluorescent Sensing Platform for the Selective and Sensitive Detection of Glyphosate. Int. J. Mol. Sci. 2021, 22, 9816. https://doi.org/10.3390/ijms22189816

AMA Style

Sun F, Ye X-L, Wang Y-B, Yue M-L, Li P, Yang L, Liu Y-L, Fu Y. NPA-Cu2+ Complex as a Fluorescent Sensing Platform for the Selective and Sensitive Detection of Glyphosate. International Journal of Molecular Sciences. 2021; 22(18):9816. https://doi.org/10.3390/ijms22189816

Chicago/Turabian Style

Sun, Fang, Xin-Lu Ye, Yu-Bo Wang, Ming-Li Yue, Ping Li, Liu Yang, Yu-Long Liu, and Ying Fu. 2021. "NPA-Cu2+ Complex as a Fluorescent Sensing Platform for the Selective and Sensitive Detection of Glyphosate" International Journal of Molecular Sciences 22, no. 18: 9816. https://doi.org/10.3390/ijms22189816

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