Next Article in Journal
Finite Element Model Updating of RC Bridge Structure with Static Load Testing: A Case Study of Vietnamese ThiThac Bridge in Coastal and Marine Environment
Next Article in Special Issue
Development of a New Voltammetric Method for Aceclofenac Determination on Glassy Carbon Electrode Modified with Hierarchical Nanocomposite
Previous Article in Journal
Research on the Derated Power Data Identification Method of a Wind Turbine Based on a Multi-Gaussian–Discrete Joint Probability Model
Previous Article in Special Issue
Development and Analysis of Electrochemical Reactor with Vibrating Functional Element for AAO Nanoporous Membranes Fabrication
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Tuning the Surface Functionality of Fe3O4 for Sensitive and Selective Detection of Heavy Metal Ions

by
Manjunatha Kumara K. S.
1,
D. H. Nagaraju
1,*,
Zhoveta Yhobu
2,
Nayan Kumar H. N.
2,
Srinivasa Budagumpi
2,
Shubhankar Kumar Bose
2,
Shivakumar P.
1 and
Venkata Narayana Palakollu
1
1
School of Applied Science, REVA University, Bengaluru 560064, India
2
Center for Nano and Material Sciences, Jain University, Jain Global Campus, Kanakapura, Ramanagaram, Bangalore 562112, India
*
Author to whom correspondence should be addressed.
Sensors 2022, 22(22), 8895; https://doi.org/10.3390/s22228895
Submission received: 14 October 2022 / Revised: 8 November 2022 / Accepted: 14 November 2022 / Published: 17 November 2022
(This article belongs to the Special Issue The Design, Fabrication and Sensor Applications of Nano-Electrodes)

Abstract

:
The functionalization of materials for ultrasensitive detection of heavy metal ions (HMIs) in the environment is crucial. Herewith, we have functionalized inexpensive and environmentally friendly Fe3O4 nanoparticles with D-valine (Fe3O4–D–Val) by a simple co-precipitation synthetic approach characterized by XRD, FE-SEM, and FTIR spectroscopy. The Fe3O4–D–Val sensor was used for the ultrasensitive detection of Cd+2, Pb+2, and Cu+2 in water samples. This sensor shows a very low detection limit of 11.29, 4.59, and 20.07 nM for Cd+2, Pb+2, and Cu+2, respectively. The detection limits are much lower than the values suggested by the world health Organization. The real water samples were also analyzed using the developed sensor.

1. Introduction

HMIs present in the environment are toxic and cause severe consequences. In particular, lead is very dangerous and causes a sweeping extent of physiological and behavioral dysfunction in animals and individuals [1]. Cadmium poisoning results in liver and kidney failure, and itai-itai disease [2]. Copper poising causes gastrointestinal and homeostasis disorders in humans [3]. HMIs poisoning is dangerous in long-term exposure and results in chronic illness and can be fatal. Therefore, selective and sensitive detection of HMIs is the foremost task in determining and monitoring contaminated water before supplying to the public. Among many techniques, the electrochemical detection method is highly advantageous over conventional techniques [4], such as inductively coupled plasma spectrometry [5], optical method of detection [6], atomic absorption spectroscopy [7], and surface plasma resonance spectroscopy [8]. Numerous studies have been focused to detect HMIs in water using a variety of nanomaterials, such as metal and metal oxide nanoparticles [9], polymeric nanomaterials [10], silicon [11], and carbon-based nanomaterials [12,13] to design nanosensors for the detection of HMIs. All these materials have limitations, either instability, lack of sensitivity, or difficulty with synthesis. Very recently, there has been an interest in metal oxide-based materials for sensing applications due to their ease of synthesis and higher stability than other materials [14]. Different metal oxide [15,16] electrodes are explored for electrochemical detection of HMIs [17]. For example, Wang et al. synthesized a porous graphitic carbon nitride/CoMn2O4 for the detection of HMIs, which displays a limit of detection (LOD) of 0.014 μM and 0.021 μM for Pb+2 and Cd+2, respectively [18]. Wei et al. prepared an effective electrochemical sensor by α-Fe2O3/NiO on glassy carbon (GC) electrode with a detection limit of 0.05, 0.08, 0.06, and 0.02 μM for Hg+2, Cd+2, Cu+2, and Pb+2, respectively [19]. Fan et al. synthesized ZnFe2O4 nanoparticles (ZFO) for the detection of HMIs and glucose by hydrothermal method. The reported ZFO-modified electrode exhibits excellent sensitivity, and LOD was found to be 7.38, 1.161, and 12.03 nM for Pb+2, Hg+2, and Cu+2, respectively [20].
Among many metal oxides, iron oxide is highly suitable for electrochemical detection of HMIs due to its eco-friendliness, minimal toxicity, and low cost, along with biocompatibility. Out of numerous techniques to synthesize iron oxides, the co-precipitation method provides scalability, and hence, easy commercialization [21]. Pure iron oxide materials have a strong adsorption affinity toward HMIs [22]. Functionalization employing organic compounds [23,24,25] enhances the adsorption potential as more active moieties can be accommodated, which helps in the uptake of HMIs. In addition, functionalization prevents agglomeration and flocculation in iron oxide [26]. Functionalization of nanomaterials also aids in the sensitive detection of HMIs. It helps in chelating the HMIs from water and thus improves the accessibility to the electrode surface. Hence, very low detection limits could be achieved [27] (Figure 1).
The square wave anodic stripping voltammetry (SWASV) is highly beneficial for acquiring the peak current at high sweep rates with better sensitivity. SWASV shrinks the background noise compared to other voltametric methods and hence lower detection limits can be achieved [28]. Typically, sensing of HMIs by the SWASV method involves the adsorption of ions by electroreduction onto the surface followed by stripping into the electrolyte. The subsequent stripping current is being estimated which corresponds to the concentration of the HMIs present in the solution. To enhance the adsorption efficiency onto the electrodes, functional groups would help in chelating the metal ions from the solution which may aid in the maximum number of ions reduced on the surface of the electrode. During the process of functionalization, D-Valine was adsorbed on the surface of Fe3O4 nanoparticles. The high sensing ability of the sensor for HMIs could be due to the availability of free NH2 groups which chelates metal ions, and hence the adsorption and reduction process would be facilitated.
It persists that the modification of nanoparticles with small or little big organic molecules (molecules with -OH, -SH, or -NH units) will move through two different recognized mechanisms for the detection of HIMs, namely, cavity entrapment and chemical affinity (or both simultaneously) [29].
We evinced a simple, single-step pattern for easy detection of HIMs. Our approach holds the advantage of affinity between Fe3O4–D–Val (having free NH2 groups) and HMIs. The interface between the NH2 group and HIMs might follow the mechanism as shown in Figure 2a. The Fe3O4–D–Val sensor was employed for the ultrasensitive and simultaneous detection of Cd+2, Pb+2, and Cu+2. These nanoparticles accomplished a very low LOD of 4.59, 11.29, and 20.07 nM for Pb+2, Cd+2, and Cu+2, respectively. The limit of detection is considerably less than the reported values and the recommended world health organization standards. The real water samples were investigated by employing a developed sensor.

2. Materials and Methods

2.1. Instrumentation

All electrochemical measurements were performed by EG&G potentiostat/galvanostat (Model 263A) in a standard three-electrode system, saturated calomel electrode (SCE) as a reference electrode, GC as a working electrode, and Pt wire as a counter electrode. The synthesized Fe3O4–D–Val was analyzed by using an infrared spectrum recorded using FTIR (Bruker, ALPHA, Billerica, MA, USA). The XRD (Rigaku X-ray diffraction Ultima-IV) was performed to study the crystal structure of synthesized material. FE-SEM (JEOL model-JSM7100F) was carried out for the morphology of Fe3O4–D–Val.

2.2. Chemicals and Materials

Ferrous sulfate (crystalline) and ferric chloride anhydrous were purchased from S. D. Fine-Chem Limited (Maharashtra, India). D-valine (99+%) was purchased from Chem-Impex international (Wood Dale, IL, USA). Liquor ammonia (About 25% NH3) was purchased from Fisher scientific (Waltham, MA, USA). Sodium acetate buffer solution (pH 5.2 ± 0.1) was acquired from Sigma Aldrich (St. Louis, MO, USA). Cadmium nitrate tetrahydrate was obtained from LOBAchemie (Maharashtra, India). Lead nitrate was purchased from S. D. Fine-Chem Limited. Cupric sulphate pentahydrate was acquired from S. D. Fine-Chem Limited. All chemicals were utilized without any further purification. Deionized water was used for all experiments.

2.3. Synthesis of Functionalized Fe3O4

Fe3O4–D–Val was synthesized by the co-precipitation method as shown in Figure 2b. Typically, 2.1 g of FeSO4.7H2O and 3.1 g of FeCl3 were dissolved in 100 mL of deionized water, and the solution was heated at 60 °C (reaction mixture). This was followed by the preparation of two solutions, i.e., (I) 10 mL of 25% liquor ammonia was dissolved in 50 mL of deionized water, (II) 0.5 g of D-valine was dissolved in 50 mL of deionized water. These two solutions were added quickly and sequentially into the solution containing the reaction mixture. The reaction mixture was kept at constant stirring with constant heating at 60 °C for about 1 h. The bright brown precipitate formed was collected after filtering followed by washing with deionized water and dried at 60 °C overnight.

2.4. Electrode Modification

Before electrode modification, the glassy carbon electrode (GCE) was cleaned with alumina slurry of various particle sizes (1, 0.5, and 0.25 μm average particle size) and subsequently washed with deionized water. The electrode was sonicated in water and ethanol (1:1) solution for about 30 min and washed with ethanol and dried at room temperature [30]. In the sensing system, GCE, with advantages of high cleanliness, high conductivity, minimal thermal extinction coefficient and wide potential window of operation, has been broadly utilized. In addition, the modified GC electrode provides phenomenal chemical property of immobilizing materials onto its surface, and aids insensitive and selective determination of HMIs. A total of 1 mg of the synthesized Fe3O4–D–Val was dispersed in a 1 mL mixture of water and ethanol solution (7:3) and ultra-sonicated for 30 min. A total of 10 μL of the above solution was drop-costed onto the surface of the cleaned GC and dried at room temperature.

3. Results and Discussion

Figure 3a shows FTIR spectra of Fe3O4–D–Val nanoparticles. The IR band at 549 cm−1 was attributed to the Fe-O stretching mode. Free amine group was confirmed by vibrational bands at 881 cm−1 attributed to the N-H wagging [31], C-C-N stretching at 1136 cm−1 and N-H bending at 1587 cm−1. The vibrational bands at 3110 and 2982 cm−1 correspond to the presence of O-H and C-H asymmetric stretching [32], respectively, and peak broadening confirms the presence of hydrogen (functional groups such as -OH, -CH3) in the synthesized nanoparticles. In the FTIR spectrum of D-valine functionalized Fe3O4 band at 1511 cm−1 is attributed to the N-H symmetric deformation [33].
The XRD technique (Figure 3b) was utilized to ascertain the crystalline structure of Fe3O4–D–Val. The XRD peaks with 2θ at 30.2°, 35.7°, 43.4°, 54.1°, 57.5°, 62.8°, and 71.9° exhibited consistency with reported data [34,35] (COD 96-900-2320) and were indexed to a cubic phase of Fe3O4 (magnetite, chemical formula: Fe24.00O32.00 Space group: Fd-3m). The typical crystallite size was found to be 8.62 nm.
The surface morphology of the synthesized nanoparticles is analyzed by FE-SEM; the composites consist of uniform particle distribution, and the bright nanoparticles arecovered by a carbon chain-like structure of amino acids, which can be observed in Figure 3c.

3.1. Electrochemical Detection of HMIs

SWASV was employed to study the electrochemical sensing of HMIs with Fe3O4–D–Val in sodium acetate buffer (0.1 M, pH 5.2 ± 0.1). Optimization of deposition potential, pH of electrolyte, and deposition time were carried out. After the optimization of experimental parameters, individual metal ion detection was carried out. Later, simultaneous detection of HIMs at optimized conditions was performed to analyze the behavior of HMIs in the presence of multiple electroactive analytes.

3.2. Influence of Deposition Potential

The anodic stripping analysis reveals that the suitable deposition potential appears to be key to achieving the highest sensitivity. The stimulus of deposition potential on the responsive stripping peak current of Pb+2 was studied by using SWASV at different potential ranges from −0.8 to −1.2 V vs. SCE in 3 μM concentration of Pb+2 for about 300 s of deposition time in sodium acetate buffer electrolyte (0.1 M and 5.2 pH) as presented in Figure 4a,b. The peak current increases with increasing deposition potential from −0.8 to −1.1 V. Deposition potential beyond −1.1 V resulted in the current decreasing. The decrease in the peak current beyond −1.1 V may be partially due to the hydrogen evolution observed at the surface of the electrode. The adsorption of the hydrogen ion intermediates on the surface of the electrode results in fewer active sites for adsorption and electro–reduction in the HMIs, thereby resulting in a reduced peak current [36].

3.3. Influence of pH of the Electrolyte

Figure 4c,d, represents the peak current of sensing Pb+2 and its relationship with pH; the increase in the pH results in the peak current increasing up to 5.2 pH and further expansion in pH prompts a decline in the peak current. At very low pH, protons compete with HMIs to bind to the electrode surface. As pH increases, the availability of protons decreases, and hence adsorption HMIs increases. At higher pH, HMIs undergo hydrolysis, which results in the peak current decreasing [37,38].

3.4. Influence of Deposition Time

The effect of the deposition time on the sensing ability was studied (Figure 4e,f); as depicted there is an increase in the peak current when the deposition current is increased from 60 to 300 s with an increment of 60 s in each successive step. The optimized deposition time as observed from the figure is 300 s and this parameter is used for further analysis irrespective of the HMIs.

3.5. Individual Detection of HMIs (Pb+2, Cd+2, and Cu+2) by SWASV

To detect the HMIs, SWASV was applied to study the GC-modified Fe3O4–D–Val as HMIs sensor in 0.1 M sodium acetate buffer electrolyte under optimized conditions. The detection of Pb+2 was studied by increasing the concentration of Pb+2 from 0.08 to 2 μM as shown in Figure 5a,b. Linear increase in anodic peak current response with increasing concentration of Pb+2 is obvious. The oxidation peak for detection of Pb+2 was at −0.51 V vs. SCE. We have observed two linear concentration ranges. At low concentration, Figure 5a and high concentration Figure 5b show linear equation Y = 0.090 X + 2.099 and Y = 0.019 X + 20.458, respectively. The LOD and sensitivity were calculated using the formula LOD = 3S/b (where S corresponds to standard deviation and b corresponds to slope from the calibration curve) and sensitivity = b/A (where A represents the area of GCE).
In addition, SWASV was applied to examine the detection of other metal ions such as cadmium (Cd+2) and copper (Cu+2), at concentrations from 0.05 to 0.8 μM and 0.1 to 1 μM, respectively. The oxidation peaks for the detection of Cd+2 and Cu+2 was observed at −0.67 and −0.07 V vs. SCE, respectively. The corresponding calibration plot was plotted as shown in Figure 5c,d. The linear equations for Cd+2 and Cu+2 are Y = 0.036 X − 0.728 and Y = 0.020 X + 6.728, respectively.

3.6. Simultaneous Detection of the HMIs by SWASV

Simultaneous detection was conducted by increasing the concentration of Cd+2, Pb+2, and Cu+2. Oxidative stripping peak was observed at −0.67 V, −0.51 V, and −0.007 V vs. SCE for Cd+2, Pb+2, and Cu+2, respectively, as displayed in Figure 6a,b. The oxidative stripping peak current increases with increasing concentration. The interpeak spacing is adequate to distinguish between the different metal ions. The anodic spacing between Cd+2 and Pb+2 is about −0.1675 V vs. SCE and −0.4469 V vs. SCE between Pb+2 and Cu+2. The LODs for Pb+2, Cd+2, and Cu+2 by simultaneous detection are 18.89, 18.38, and 7.48 nM, respectively. Comparison of different electrode material LODs are listed in Table 1. The LODs for simultaneous and individual HIMs detection of as synthesized electrode material with their thresh hold limits by WHO is shown in Table 2. Compared to individual HMIs detection, the results appear to rise in LOD values (Pb+2 by 14.33 nM and Cd+2 by 7.09 nM). Cu+2 showed comparatively low LOD (Cu+2 by 12.609 nM), which could be ascribed to the interface between different HMIs, such as the development of intermetallic compounds and competitive deposition [39]. The outcomes from simultaneous detection point to the application of developed sensors for real-time water analysis.

3.7. Reproducibility, Stability and Interference Study’s

Fe3O4–D–Val/GC-modified electrode exhibits electrochemical detection towards Cd+2, Pb+2, and Cu+2. To evaluate the reproducibility of the sensor, SWASV was employed to detect 1 μM Pb+2, 2 μM Cd+2, and 2 μM Cu+2 in sodium acetate buffer (0.1 M) for ten separate trials. The relative standard deviation (RSD) is 2.37%, 7.94% and 11.54%, for Pb+2, Cd+2 and Cu+2 ions, respectively. The stability of the electrode was also analyzed under similar conditions, and the anodic stripping current response remains at 78.04%, 85.10%, and 73.66% even after 16 days as shown in the Figure 6c,d. Further, to evaluate the interference of other metal ions on the as-prepared sensor, 1 μM Pb2+ was followed by 0.1 μM of conceivable interfering metal ions such as Cd2+, Fe2+, Cr3+, Hg2+, Ni2+, and Cu2+. It was found that there is no prevalent change in ht epeak current of Pb2+ ((M) = 1 μM) as depicted in Figure 6f.

3.8. Real-Time Applications Study

As per the world health organization, the permissible concentration of Pb2+ in drinking or river water is 10 μg/L. As depicted in Figure 5a,b, the Fe3O4–D–Val electrode allows for detection as low as WHO permissible limits. On this basis, Fe3O4–D–Val electrodes were employed for the analysis of real water samples (River water: 13°45′11.6″N 76°53′45.3″, 200F and Time 2:20 PM). The real sample shows no peak corresponding to the heavy metal ion. To confirm this further, the actual sample was spiked with Pb2+ ions. The samples were spiked in a 3:7 volume ratio of real water sample and 0.1 M sodium acetate buffer [51]. The spiked sample displays a peak corresponding to the Pb2+ ions. However, the peak current in real sample water is less than in the clean water when measuring individual Pb+2 ions, as is evident in Figure 6e. This could be due to the influence of other natural impurities and industrial effluents in a real sample as well as a lack of ionic species in the electrolyte, thereby reducing the ionic conductivity in the electrolyte and resulting in a lesser anodic peak current.

4. Conclusions

An easy, cost-efficient, and sustainable synthesis pathway for the preparation of iron NPs-integrated D-Valine materials has been developed and successfully employed for SWASV detection of HMIs, i.e., Cd+2, Pb+2, and Cu+2. The Fe3O4–D–Val-modified GCE was found to exhibit excellent selectivity, sensitivity, and low LOD for selective as well as simultaneous detection of HMIs. The modified electrode was also found to detect HMIs in the real water sample. This positive result in this work evokes the use of Fe3O4–D–Val/GCE as a sensing platform for the detection of HMIs in the water.

Author Contributions

Manuscript—original draft preparation, M.K.K.S.; methodology, M.K.K.S.; Conceptualization, M.K.K.S.; investigation, D.H.N.; manuscript correction, S.B. and S.K.B.; supervision, Z.Y., N.K.H.N., S.P. and V.N.P.; funding acquisition, D.H.N. All authors have read and agreed to the published version of the manuscript.

Funding

We are appreciative of the financial support from the Department of Science and Technology, Technology Development Transfer Division under the Device Development Program (DDP) (Grant No. DST/TDT/DDP-33/2018). This work was also partially supported by SEED money from REVA university (RU:EST:CHE:2021/22).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors proclaim that they have not acknowledged opposing financial aids or individual interactions that could have seemed to impact the work informed in this paper.

References

  1. Hsu, P.C.; Guo, Y.L. Antioxidant Nutrients and Lead Toxicity. Toxicology 2002, 180, 33–44. [Google Scholar] [CrossRef]
  2. Inaba, T.; Kobayashi, E.; Suwazono, Y.; Uetani, M.; Oishi, M.; Nakagawa, H.; Nogawa, K. Estimation of Cumulative Cadmium Intake Causing Itai-Itai Disease. Toxicol. Lett. 2005, 159, 192–201. [Google Scholar] [CrossRef] [PubMed]
  3. Taylor, A.A.; Tsuji, J.S.; Garry, M.R.; McArdle, M.E.; Goodfellow, W.L.; Adams, W.J.; Menzie, C.A. Critical Review of Exposure and Effects: Implications for Setting Regulatory Health Criteria for Ingested Copper. Environ. Manag. 2020, 65, 131–159. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Malik, L.A.; Bashir, A.; Qureashi, A.; Pandith, A.H. Detection and Removal of Heavy Metal Ions: A Review. Environ. Chem. Lett. 2019, 17, 1495–1521. [Google Scholar] [CrossRef]
  5. Griffiths, A.; Packman, H.; Leung, Y.L.; Coles, B.J.; Kreissig, K.; Little, S.H.; Van De Flierdt, T.; Rehkämper, M. Evaluation of Optimized Procedures for High-Precision Lead Isotope Analyses of Seawater by Multiple Collector Inductively Coupled Plasma Mass Spectrometry. Anal. Chem. 2020, 92, 11232–11241. [Google Scholar] [CrossRef]
  6. Na Kim, H.; Xiu Ren, W.; Seung Kim, J.; Yoon, J. Fluorescent and Colorimetric Sensors for Detection of Lead, Cadmium, and Mercury Ions. Chem. Soc. Rev. 2012, 41, 3210–3244. [Google Scholar] [CrossRef]
  7. Altunay, N.; Elik, A.; Bingöl, D. Simple and Green Heat-Induced Deep Eutectic Solvent Microextraction for Determination of Lead and Cadmium in Vegetable Samples by Flame Atomic Absorption Spectrometry: A Multivariate Study. Biol. Trace Elem. Res. 2020, 198, 324–331. [Google Scholar] [CrossRef]
  8. Forzani, E.S.; Zhang, H.; Chen, W.; Tao, N. Detection of Heavy Metal Ions in Drinking Water Using a High-Resolution Differential Surface Plasmon Resonance Sensor. Environ. Sci. Technol. 2005, 39, 1257–1262. [Google Scholar] [CrossRef]
  9. Sawan, S.; Maalouf, R.; Errachid, A.; Jaffrezic-Renault, N. Metal and Metal Oxide Nanoparticles in the Voltammetric Detection of Heavy Metals: A Review. TrAC-Trends Anal. Chem. 2020, 131, 116014. [Google Scholar] [CrossRef]
  10. Lo, M.; Ktari, N.; Gningue-Sall, D.; Madani, A.; Aaron, S.E.; Aaron, J.-J.; Mekhalif, Z.; Delhalle, J.; Chehimi, M.M. Polipirrol: Polímero Conductor Reactivo y Funcional Para La Detección Electroquímica Selectiva de Metales Pesados En Agua. Emergent Mater. 2020, 3, 815–839. [Google Scholar] [CrossRef]
  11. Yin, J.; Gao, W.; Zhang, Z.; Mai, Y.; Luan, A.; Jin, H.; Jian, J.; Jin, Q. Batch Microfabrication of Highly Integrated Silicon-Based Electrochemical Sensor and Performance Evaluation via Nitrite Water Contaminant Determination. Electrochim. Acta 2020, 335, 135660. [Google Scholar] [CrossRef]
  12. Lochab, A.; Sharma, R.; Kumar, S.; Saxena, R. Recent Advances in Carbon Based Nanomaterials as Electrochemical Sensor for Toxic Metal Ions in Environmental Applications. Mater. Today Proc. 2020, 45, 3741–3753. [Google Scholar] [CrossRef]
  13. Pandey, S.K.; Singh, P.; Singh, J.; Sachan, S.; Srivastava, S.; Singh, S.K. Nanocarbon-Based Electrochemical Detection of Heavy Metals. Electroanalysis 2016, 28, 2472–2488. [Google Scholar] [CrossRef]
  14. Aragay, G.; Pons, J.; Merkoçi, A. Recent Trends in Macro-, Micro-, and Nanomaterial-Based Tools and Strategies for Heavy-Metal Detection. Chem. Rev. 2011, 111, 3433–3458. [Google Scholar] [CrossRef] [PubMed]
  15. Yang, M.; Li, P.H.; Chen, S.H.; Xiao, X.Y.; Tang, X.H.; Lin, C.H.; Huang, X.J.; Liu, W.Q. Nanometal Oxides with Special Surface Physicochemical Properties to Promote Electrochemical Detection of Heavy Metal Ions. Small 2020, 16, 1–19. [Google Scholar] [CrossRef]
  16. Yu, X.Y.; Meng, Q.Q.; Luo, T.; Jia, Y.; Sun, B.; Li, Q.X.; Liu, J.H.; Huang, X.J. Facet-Dependent Electrochemical Properties of Co 3 O 4 Nanocrystals toward Heavy Metal Ions. Sci. Rep. 2013, 3, 1–7. [Google Scholar] [CrossRef] [Green Version]
  17. Wei, Y.; Gao, C.; Meng, F.L.; Li, H.H.; Wang, L.; Liu, J.H.; Huang, X.J. SnO2/Reduced Graphene Oxide Nanocomposite for the Simultaneous Electrochemical Detection of Cadmium(II), Lead(II), Copper(II), and Mercury(II): An Interesting Favorable Mutual Interference. J. Phys. Chem. C 2012, 116, 1034–1041. [Google Scholar] [CrossRef]
  18. Wang, Y.; Nie, Z.; Li, X.; Zhao, Y.; Wang, H. Highly Sensitive and Selective Electrochemical Sensor Based on Porous Graphitic Carbon Nitride/CoMn2O4 Nanocomposite toward Heavy Metal Ions. Sens. Actuators B Chem. 2021, 346, 130539. [Google Scholar] [CrossRef]
  19. Wei, J.; Zhao, J.; Li, C.Y.; Xie, X.Y.; Wei, Y.Y.; Shen, W.; Wang, J.P.; Yang, M. Highly Sensitive and Selective Electrochemical Detection of Pb(II) in Serum via an α-Fe2O3/NiO Heterostructure: Evidence from Theoretical Calculations and Adsorption Investigation. Sens. Actuators B Chem. 2021, 344, 130295. [Google Scholar] [CrossRef]
  20. Fan, C.; Chen, L.; Jiang, R.; Ye, J.; Li, H.; Shi, Y.; Luo, Y.; Wang, G.; Hou, J.; Guo, X. ZnFe2O4Nanoparticles for Electrochemical Determination of Trace Hg(II), Pb(II), Cu(II), and Glucose. ACS Appl. Nano Mater. 2021, 4, 4026–4036. [Google Scholar] [CrossRef]
  21. Besenhard, M.O.; LaGrow, A.P.; Hodzic, A.; Kriechbaum, M.; Panariello, L.; Bais, G.; Loizou, K.; Damilos, S.; Margarida Cruz, M.; Thanh, N.T.K.; et al. Co-Precipitation Synthesis of Stable Iron Oxide Nanoparticles with NaOH: New Insights and Continuous Production via Flow Chemistry. Chem. Eng. J. 2020, 399, 125740. [Google Scholar] [CrossRef]
  22. Ramos Guivar, J.A.; Sadrollahi, E.; Menzel, D.; Ramos Fernandes, E.G.; López, E.O.; Torres, M.M.; Arsuaga, J.M.; Arencibia, A.; Litterst, F.J. Magnetic, Structural and Surface Properties of Functionalized Maghemite Nanoparticles for Copper and Lead Adsorption. RSC Adv. 2017, 7, 28763–28779. [Google Scholar] [CrossRef] [Green Version]
  23. Mohapatra, S.; Pramanik, P. Synthesis and Stability of Functionalized Iron Oxide Nanoparticles Using Organophosphorus Coupling Agents. Colloids Surf. A Physicochem. Eng. Asp. 2009, 339, 35–42. [Google Scholar] [CrossRef]
  24. Mai, T.; Hilt, J.Z. Functionalization of Iron Oxide Nanoparticles with Small Molecules and the Impact on Reactive Oxygen Species Generation for Potential Cancer Therapy. Colloids Surf. A Physicochem. Eng. Asp. 2019, 576, 9–14. [Google Scholar] [CrossRef]
  25. Mikhaylova, M.; Kim, D.K.; Berry, C.C.; Zagorodni, A.; Toprak, M.; Curtis, A.S.G.; Muhammed, M. BSA Immobilization on Amine-Functionalized Superparamagnetic Iron Oxide Nanoparticles. Chem. Mater. 2004, 16, 2344–2354. [Google Scholar] [CrossRef]
  26. Wu, W.; He, Q.; Jiang, C. Magnetic Iron Oxide Nanoparticles: Synthesis and Surface Functionalization Strategies. Nanoscale Res. Lett. 2008, 3, 397–415. [Google Scholar] [CrossRef] [Green Version]
  27. Jijana, A.N.; Mphuthi, N.; Shumbula, P.; Vilakazi, S.; Sikhwivhilu, L. The Ultra-Sensitive Electrochemical Detection of As(III) in Ground Water Using Disposable L-Cysteine/Lipoic Acid Functionalised Gold Nanoparticle Modified Screen-Printed Electrodes. Electrocatalysis 2021, 12, 310–325. [Google Scholar] [CrossRef]
  28. Arantes, T.M.; Sardinha, A.; Baldan, M.R.; Cristovan, F.H.; Ferreira, N.G. Lead Detection Using Micro/Nanocrystalline Boron-Doped Diamond by Square-Wave Anodic Stripping Voltammetry. Talanta 2014, 128, 132–140. [Google Scholar] [CrossRef]
  29. Zuo, Y.; Xu, J.; Zhu, X.; Duan, X.; Lu, L.; Yu, Y. Graphene-Derived Nanomaterials as Recognition Elements for Electrochemical Determination of Heavy Metal Ions: A Review. Microchim. Acta 2019, 186, 1–17. [Google Scholar] [CrossRef]
  30. Muralikrishna, S.; Sureshkumar, K.; Varley, T.S.; Nagaraju, D.H.; Ramakrishnappa, T. Analytical Methods In Situ Reduction and Functionalization of Graphene Oxide with L -Cysteine for Simultaneous Electrochemical Determination of Cadmium (II). Anal. Methods 2014, 6, 8698–8705. [Google Scholar] [CrossRef]
  31. Sun, Y.F.; Zhao, L.J.; Jiang, T.J.; Li, S.S.; Yang, M.; Huang, X.J. Sensitive and Selective Electrochemical Detection of Heavy Metal Ions Using Amino-Functionalized Carbon Microspheres. J. Electroanal. Chem. 2016, 760, 143–150. [Google Scholar] [CrossRef]
  32. Ting, S.L.; Ee, S.J.; Ananthanarayanan, A.; Leong, K.C.; Chen, P. Graphene Quantum Dots Functionalized Gold Nanoparticles for Sensitive Electrochemical Detection of Heavy Metal Ions. Electrochim. Acta 2015, 172, 7–11. [Google Scholar] [CrossRef]
  33. Akbar, A.; Riaz, S.; Ashraf, R.; Naseem, S. Magnetic and Magnetization Properties of Iron Oxide Thin Films by Microwave Assisted Sol–Gel Route. J. Sol-Gel Sci. Technol. 2015, 74, 320–328. [Google Scholar] [CrossRef]
  34. Haavik, C.; Stølen, S.; Fjellvåg, H.; Hanfland, M.; Häusermann, D. Equation of State of Magnetite and Its High-Pressure Modification: Thermodynamics of the Fe-O System at High Pressure. Am. Mineral. 2000, 85, 514–523. [Google Scholar] [CrossRef]
  35. Rahman, S.S.U.; Qureshi, M.T.; Sultana, K.; Rehman, W.; Khan, M.Y.; Asif, M.H.; Farooq, M.; Sultana, N. Single Step Growth of Iron Oxide Nanoparticles and Their Use as Glucose Biosensor. Results Phys. 2017, 7, 4451–4456. [Google Scholar] [CrossRef]
  36. Muralikrishna, S.; Nagaraju, D.H.; Balakrishna, R.G.; Surareungchai, W.; Ramakrishnappa, T.; Shivanandareddy, A.B. Hydrogels of Polyaniline with Graphene Oxide for Highly Sensitive Electrochemical Determination of Lead Ions. Anal. Chim. Acta 2017, 990, 67–77. [Google Scholar] [CrossRef]
  37. Suvina, V.; Krishna, S.M.; Nagaraju, D.H.; Melo, J.S.; Balakrishna, R.G. Polypyrrole-Reduced Graphene Oxide Nanocomposite Hydrogels: A Promising Electrode Material for the Simultaneous Detection of Multiple Heavy Metal Ions. Mater. Lett. 2018, 232, 209–212. [Google Scholar] [CrossRef]
  38. Afkhami, A.; Soltani-Felehgari, F.; Madrakian, T.; Ghaedi, H.; Rezaeivala, M. Fabrication and Application of a New Modified Electrochemical Sensor Using Nano-Silica and a Newly Synthesized Schiff Base for Simultaneous Determination of Cd2+, Cu2+ and Hg2+ Ions in Water and Some Foodstuff Samples. Anal. Chim. Acta 2013, 771, 21–30. [Google Scholar] [CrossRef]
  39. Lu, M.; Deng, Y.; Luo, Y.; Lv, J.; Li, T.; Xu, J.; Chen, S.W.; Wang, J. Graphene Aerogel-Metal-Organic Framework-Based Electrochemical Method for Simultaneous Detection of Multiple Heavy-Metal Ions. Anal. Chem. 2019, 91, 888–895. [Google Scholar] [CrossRef]
  40. Munir, A.; Shah, A.; Nisar, J.; Ashiq, M.N.; Akhter, M.S.; Shah, A.H. Selective and Simultaneous Detection of Zn2+, Cd2+, Pb2+, Cu2+, Hg2+ and Sr2+ Using Surfactant Modified Electrochemical Sensors. Electrochim. Acta 2019, 323, 134592. [Google Scholar] [CrossRef]
  41. Zhu, F.; Shi, H.; Yu, Z.; Wang, C.; Cheng, W.; Zhou, X.; Yang, F.; Zhang, Y.; Zhang, X. Acid-Etched Fe/Fe2O3 nanoparticles Encapsulated into Carbon Cloth as a Novel Voltammetric Sensor for the Simultaneous Detection of Cd2+ and Pb2. Analyst 2021, 146, 691–697. [Google Scholar] [CrossRef] [PubMed]
  42. Yamuna, A.; Hong, C.Y.; Chen, S.M.; Chen, T.W.; Alabdullkarem, E.A.; Soylak, M.; Mana AL-Anazy, M.; Ajmal Ali, M.; Liu, X. Highly Selective Simultaneous Electrochemical Detection of Trace Level of Heavy Metals in Water Samples Based on the Single-Crystalline Co3O4 Nanocubes Modified Electrode. J. Electroanal. Chem. 2021, 887, 115159. [Google Scholar] [CrossRef]
  43. Ru, J.; Wang, X.; Cui, X.; Wang, F.; Ji, H.; Du, X.; Lu, X. GaOOH-Modified Metal-Organic Frameworks UiO-66-NH2: Selective and Sensitive Sensing Four Heavy-Metal Ions in Real Wastewater by Electrochemical Method. Talanta 2021, 234, 122679. [Google Scholar] [CrossRef] [PubMed]
  44. Ding, Y.; Wei, F.; Dong, C.; Li, J.; Zhang, C.; Han, X. UiO-66 Based Electrochemical Sensor for Simultaneous Detection of Cd(II) and Pb(II). Inorg. Chem. Commun. 2021, 131, 108785. [Google Scholar] [CrossRef]
  45. Sha, H.; Wu, Y.; Fan, Y. A Fe-OSA/Nafion Composite Film-Decorated Glassy Carbon Electrode as a Sensor for Detection of Pb(II), Cd(II) and Cu(II). Anal. Methods 2017, 9, 5618–5631. [Google Scholar] [CrossRef]
  46. Pei, J.; Yu, X.; Wei, S.; Boukherroub, R.; Zhang, Y. Double-Side Effect of B/C Ratio on BDD Electrode Detection for Heavy Metal Ion in Water. Sci. Total Environ. 2021, 771, 145430. [Google Scholar] [CrossRef]
  47. Wang, J.; Yu, P.; Kan, K.; Lv, H.; Liu, Z.; Sun, B.; Bai, X.; Chen, J.; Zhang, Y.; Shi, K. Efficient Ultra-Trace Electrochemical Detection of Cd2+, Pb2+ and Hg2+ Based on Hierarchical Porous S-Doped C3N4 Tube Bundles/Graphene Nanosheets Composite. Chem. Eng. J. 2021, 420, 130317. [Google Scholar] [CrossRef]
  48. Veerakumar, P.; Veeramani, V.; Chen, S.M.; Madhu, R.; Liu, S. Bin Palladium Nanoparticle Incorporated Porous Activated Carbon: Electrochemical Detection of Toxic Metal Ions. ACS Appl. Mater. Interfaces 2016, 8, 1319–1326. [Google Scholar] [CrossRef]
  49. Xu, R.X.; Yu, X.Y.; Gao, C.; Liu, J.H.; Compton, R.G.; Huang, X.J. Enhancing Selectivity in Stripping Voltammetry by Different Adsorption Behaviors: The Use of Nanostructured Mg-Al-Layered Double Hydroxides to Detect Cd(Ii). Analyst 2013, 138, 1812–1818. [Google Scholar] [CrossRef]
  50. Flores, E.; Pizarro, J.; Godoy, F.; Segura, R.; Gómez, A.; Agurto, N.; Sepúlveda, P. An Electrochemical Sensor for the Determination of Cu(II) Using a Modified Electrode with Ferrocenyl Crown Ether Compound by Square Wave Anodic Stripping Voltammetry. Sens. Actuators B Chem. 2017, 251, 433–439. [Google Scholar] [CrossRef]
  51. Xu, R.X.; Yu, X.Y.; Gao, C.; Jiang, Y.J.; Han, D.D.; Liu, J.H.; Huang, X.J. Non-Conductive Nanomaterial Enhanced Electrochemical Response in Stripping Voltammetry: The Use of Nanostructured Magnesium Silicate Hollow Spheres for Heavy Metal Ions Detection. Anal. Chim. Acta 2013, 790, 31–38. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Advantages of functionalized Fe3O4.
Figure 1. Advantages of functionalized Fe3O4.
Sensors 22 08895 g001
Figure 2. (a) A possible mechanism of interaction of HIMs with free NH2 groups of Fe3O4–D–Val, (I) Fe3O4–D–Val modified glassy carbon electrode, and (II) Interaction of HMIs with NH2 groups of Fe3O4–D–Val. (b) Schematic synthesis of Fe3O4–D–Val by co-precipitation process.
Figure 2. (a) A possible mechanism of interaction of HIMs with free NH2 groups of Fe3O4–D–Val, (I) Fe3O4–D–Val modified glassy carbon electrode, and (II) Interaction of HMIs with NH2 groups of Fe3O4–D–Val. (b) Schematic synthesis of Fe3O4–D–Val by co-precipitation process.
Sensors 22 08895 g002
Figure 3. (a) FT-IR spectra of Fe3O4–D–Val. (b) XRD pattern of Fe3O4–D–Val. (c) FE-SEM images of Fe3O4–D–Val.
Figure 3. (a) FT-IR spectra of Fe3O4–D–Val. (b) XRD pattern of Fe3O4–D–Val. (c) FE-SEM images of Fe3O4–D–Val.
Sensors 22 08895 g003
Figure 4. (a) voltammogram of optimization of deposition potential, and (b) Optimization of deposition potential by applying SWASV at different potentials (300 s deposition time and (Pb+2) = 3 μM). (c) voltammogram of optimization of pH of the electrolyte, and (d) Optimization of pH of the electrolyte (optimized deposition potential: −1.1 V vs. SCE, deposition time: 300 s and (Pb+2) = 3 μM). (e) voltammogram of optimization of deposition time, and (f) Optimization of deposition time by SWASV (0.1 M sodium acetate buffer electrolyte, deposition potential −1.1 V vs. SCE and (Pb+2) = 3 μM).
Figure 4. (a) voltammogram of optimization of deposition potential, and (b) Optimization of deposition potential by applying SWASV at different potentials (300 s deposition time and (Pb+2) = 3 μM). (c) voltammogram of optimization of pH of the electrolyte, and (d) Optimization of pH of the electrolyte (optimized deposition potential: −1.1 V vs. SCE, deposition time: 300 s and (Pb+2) = 3 μM). (e) voltammogram of optimization of deposition time, and (f) Optimization of deposition time by SWASV (0.1 M sodium acetate buffer electrolyte, deposition potential −1.1 V vs. SCE and (Pb+2) = 3 μM).
Sensors 22 08895 g004
Figure 5. (a) SWASV of Pb+2 was applied on modified glassy carbon under optimized conditions at low and (b) high concentrations. (c) SWASV detection of Cd+2 and (d) SWASV detection of Cu+2 under optimized conditions. Different color line’s corresponds to the different concentration of heavy metal ions. With in graph its difficult to display all concentartions, the concentartion rage from minimal to maximum is displayed with upword arrow mark.
Figure 5. (a) SWASV of Pb+2 was applied on modified glassy carbon under optimized conditions at low and (b) high concentrations. (c) SWASV detection of Cd+2 and (d) SWASV detection of Cu+2 under optimized conditions. Different color line’s corresponds to the different concentration of heavy metal ions. With in graph its difficult to display all concentartions, the concentartion rage from minimal to maximum is displayed with upword arrow mark.
Sensors 22 08895 g005
Figure 6. (a) SWASV for simultaneous detection of three different metal ions at optimized conditions and (b) Calibration plot from simultaneous detection. (c) Anodic stripping response by SWASV for reproducibility of Fe3O4–D–Val/GC-modified electrode for detection of 1 μM Pb+2, 2 μM Cd+2, and 2 μM Cu+2 in 0.1 M sodium acetate buffer solution. (d) Stability of the modified glassy carbon electrode at a constant concentration of HMIs (2 μM). (e) SWASV peak spiking of Pb+2 at 1 μM concentration. (f) Effect of interference species. Different color line’s corresponds to the different concentration of heavy metal ions. With in graph its difficult to display all concentartions, the concentartion rage from minimal to maximum is displayed with upword arrow mark.
Figure 6. (a) SWASV for simultaneous detection of three different metal ions at optimized conditions and (b) Calibration plot from simultaneous detection. (c) Anodic stripping response by SWASV for reproducibility of Fe3O4–D–Val/GC-modified electrode for detection of 1 μM Pb+2, 2 μM Cd+2, and 2 μM Cu+2 in 0.1 M sodium acetate buffer solution. (d) Stability of the modified glassy carbon electrode at a constant concentration of HMIs (2 μM). (e) SWASV peak spiking of Pb+2 at 1 μM concentration. (f) Effect of interference species. Different color line’s corresponds to the different concentration of heavy metal ions. With in graph its difficult to display all concentartions, the concentartion rage from minimal to maximum is displayed with upword arrow mark.
Sensors 22 08895 g006
Table 1. Comparison of LOD and sensitivity of different electrodes reported in the literature.
Table 1. Comparison of LOD and sensitivity of different electrodes reported in the literature.
Modification
Material
Measurement
Technique
LODSensitivityRef
Pb+2Cd+2Cu+2
Pg-C3N4/CoMn2O4SWASV0.014 μM0.021 μM Pb+2:22.39
(μA/μM cm2)
[18]
CI-DPTU/GCESWASV11.0 nM6.45 nM7.85 nM [40]
ae-Fe/Fe2O3@ccDPASV0.5 ppb0.42 ppb Pb+2: 408.0
Cd+2: 338.7
(μA/μM cm2)
[41]
Co3O4-NC/SPCEDPV0.00722 μM 0.00173 μMPb+2: 16.73
Cu+2: 11.46
(μA/μM cm2)
[42]
UiO-66- NH2/GaOOHDPV0.028 μM0.016 μM0.019 μM [43]
UiO-66/Bi/GCESWASV0.94 µg/L2.01 µg/L [44]
Fe-OSADPV0.0360 μM0.0192 μM [45]
SBDDESWASV5–120 μg/L Pb+2: 0.42
(μA/μM cm2)
[46]
S-doped C3N4 tube bundles/graphene nanosheets compositeSWASV0.78 nM2.30 nM [47]
Pd1.5/PAC-900DPV50 nM41 nM66 nMPb+2: 109.1
Cd+2: 72.9
Cu+2: 21.8
(μA/μM cm2)
[48]
Mg–Al-LDH/NafionSWASV 0.20 nM Cd+2: 13.86 mA mM−1[49]
CB-15-crown-5-GEC,
GC/FcIB15C5
DPV,
SWASV
2.3,
0.11 g/L
[50]
D-valine functionalized Fe3O4SWASV4.59 nM11.29 nM20.07 nMPb+2: 1.275
Cd+2: 0.518
Cu+2: 0.291
(μA/nM cm2)
This work
Table 2. Comparison of detection limits obtained for Cd+2, Pb+2, and Cu+2 with the Fe3O4–D–Val sensor and WHO [30] threshold limits.
Table 2. Comparison of detection limits obtained for Cd+2, Pb+2, and Cu+2 with the Fe3O4–D–Val sensor and WHO [30] threshold limits.
Metal IonsLOD a
(nM)
Sensitivity
(μA/nM cm2)
LOD b
(nM)
Sensitivity
(μA/nM cm2)
WHO Standard Values (mg/L)
Pb+24.591.27518.890.31010.01
Cd+211.290.51818.380.31860.003
Cu+220.070.2917.4810.78322.0
a: LODs by individual ion detection, b: LODs by simultaneous detection.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

K. S., M.K.; Nagaraju, D.H.; Yhobu, Z.; H. N., N.K.; Budagumpi, S.; Kumar Bose, S.; P., S.; Palakollu, V.N. Tuning the Surface Functionality of Fe3O4 for Sensitive and Selective Detection of Heavy Metal Ions. Sensors 2022, 22, 8895. https://doi.org/10.3390/s22228895

AMA Style

K. S. MK, Nagaraju DH, Yhobu Z, H. N. NK, Budagumpi S, Kumar Bose S, P. S, Palakollu VN. Tuning the Surface Functionality of Fe3O4 for Sensitive and Selective Detection of Heavy Metal Ions. Sensors. 2022; 22(22):8895. https://doi.org/10.3390/s22228895

Chicago/Turabian Style

K. S., Manjunatha Kumara, D. H. Nagaraju, Zhoveta Yhobu, Nayan Kumar H. N., Srinivasa Budagumpi, Shubhankar Kumar Bose, Shivakumar P., and Venkata Narayana Palakollu. 2022. "Tuning the Surface Functionality of Fe3O4 for Sensitive and Selective Detection of Heavy Metal Ions" Sensors 22, no. 22: 8895. https://doi.org/10.3390/s22228895

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