Next Article in Journal
Spin-Mechanics with Nitrogen-Vacancy Centers and Trapped Particles
Previous Article in Journal
Highly Sensitive Hybrid Nanostructures for Dimethyl Methyl Phosphonate Detection
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Direct Mercury Detection in Landfill Leachate Using a Novel AuNP-Biopolymer Carbon Screen-Printed Electrode Sensor

1
Department of Civil, Environmental, and Construction Engineering, University of Central Florida, Orlando, FL 32816, USA
2
NanoScience Technology Center and Department of Chemistry, University of Central Florida, Orlando, FL 32816, USA
3
Department of Chemistry, University of Central Florida, Orlando, FL 32816, USA
*
Authors to whom correspondence should be addressed.
Micromachines 2021, 12(6), 649; https://doi.org/10.3390/mi12060649
Submission received: 5 May 2021 / Revised: 26 May 2021 / Accepted: 30 May 2021 / Published: 1 June 2021
(This article belongs to the Special Issue Microsensors for Water Monitoring)

Abstract

:
A novel Au nanoparticle (AuNP)-biopolymer coated carbon screen-printed electrode (SPE) sensor was developed through the co-electrodeposition of Au and chitosan for mercury (Hg) ion detection. This new sensor showed successful Hg2+ detection in landfill leachate using square wave anodic stripping voltammetry (SWASV) with an optimized condition: a deposition potential of −0.6 V, deposition time of 200 s, amplitude of 25 mV, frequency of 60 Hz, and square wave step voltage of 4 mV. A noticeable peak was observed at +0.58 V associated with the stripping current of the Hg ion. The sensor exhibited a good sensitivity of ~0.09 μA/μg (~0.02 μA/nM) and a linear response over the concentration range of 10 to 100 ppb (50–500 nM). The limit of detection (LOD) was 1.69 ppb, which is significantly lower than the safety limit defined by the United States Environmental Protection Agency (USEPA). The sensor had an excellent selective response to Hg2+ in landfill leachate against other interfering cations (e.g., Zn2+, Pb2+, Cd2+, and Cu2+). Fifteen successive measurements with a stable peak current and a lower relative standard deviation (RSD = 5.1%) were recorded continuously using the AuNP-biopolymer-coated carbon SPE sensor, which showed excellent stability, sensitivity and reproducibility and consistent performance in detecting the Hg2+ ion. It also exhibited a good reliability and performance in measuring heavy metals in landfill leachate.

Graphical Abstract

1. Introduction

Mercury (Hg) pollution caused by industrial activity has been attracting global attention for decades [1,2,3]. Mercury occurs in many forms in aqueous solution depending on oxidation and reduction conditions [4]. The main forms of Hg exposure in the general population include methylmercury (MeHg) from seafood, inorganic mercury (I–Hg) from food, and mercury vapor (Hg0) from dental amalgam fillings [5]. Most Hg occurs in organic and inorganic forms of divalent mercury and Hg0, as a form of Hg dissolved in an aqueous phase [6]. Once Hg has reached surface waters or soils, microorganisms convert it to MeHg, a substance that can be absorbed quickly by most organisms including marine life. MeHg, upon consumption, can cause negative health effects (i.e., nerve, kidney and intestinal damage; stomach disruption; reproductive failure; and DNA alteration) [7]. Inorganic mercury, Hg0 and Hg2+, is released into the environment from a variety of anthropogenic and natural sources. In particular, Hg2+ ions are one of the largest hazardous Hg pollutants in aquatic ecosystems [8]. Furthermore, Hg2+ is not readily biodegradable and is prone to bioaccumulation and biomagnification across trophic levels. Exposure to Hg2+ has been linked to several diseases; such as Minamata [9], acrodynia [10], cardiac [11] and neurological disorders, [12] and several developmental illnesses.
Surface water and groundwater are often polluted by untreated sewage water, industrial waste, gasoline, medical waste, electroplated steel and electronic parts. Since Hg accumulates in the ecosystem during tropospheric cycling, it is essential to regulate the presence of Hg2+ in drinking water and environmental systems. The United States Environmental Protection Agency (USEPA) has mandated an upper limit of 2 ppb (10 nM) of Hg2+ in drinking water [3], so monitoring of very low concentrations of Hg in the early stages of pollution is required to assess hidden risks. Various analytical methods such as atomic absorption spectrophotometry (AAS) [13], atomic fluorescence spectrometry (AFS) [14], inductively coupled plasma mass spectrometry (ICP-MS) [15], and X-ray fluorescence spectrometry [16] have been developed for detecting Hg. These are generally combined with cold vapor generation and amalgamation techniques to separate and pre-concentrate Hg to achieve a high sensitivity [17,18]. However, the complicated pretreatment requires a variety of different reagents and numerous preparation steps. Electrochemical techniques are very effective in detecting low concentrations of mercury because metal ions can be rapidly pre-concentrated on the electrode surface using methods such as anodic stripping voltammetry (ASV) [19]. In addition, modifying electrodes can offer ease of operation and low-cost and portable instrumentation [20]. The sensitivity and selectivity of Hg electrochemical sensors have been greatly improved by nanostructured electrodes, which provide a large surface area and intriguing properties [21,22,23].
Gold nanostructures have attracted great research interest in the voltammetric detection of mercury because of their high affinity [24], good catalytic ability [25], high adsorption capacities [26], and excellent mass transport [27,28]. Among the many methods to prepare AuNPs, the reduction of chloroauric acid (HAuCl4) is the most popular [29], but recently, the use of polymers to synthesize AuNPs has been reported. For example, AuNPs were prepared in the presence of chitosan [21,22] and its derivative carboxymethylated chitosan [24]. Chitosan is recognized as a low-cost promising supplement to AuNPs and other metal film electrodes due to its relatively high mechanical strength, good adhesion to traditional electrochemical surfaces, high water permeability, biocompatibility and [30,31] ability to form stable chelates with transition metal ions [32]. However, the application of chitosan-coated AuNPs has been limited to specific applications due to their tendency to agglomerate and precipitate; therefore coupling with a suitable supporting substrate may increase their viability in preparation of metal composite sensors [33].
In this study, we demonstrated direct Hg2+ detection in real landfill leachate samples using a AuNP-biopolymer-coated carbon SPE sensor. AuNP and the biopolymer (chitosan) were coated onto a carbon SPE sensor by electrodeposition and gave enhanced sensing performance for Hg2+ detection through an increase in both electrochemical sensitivity and stability. Using the new sensor, systematic batch experiments were performed to determine the optimal deposition time amplitude and frequency for detecting Hg2+ ions using square wave anodic stripping voltammetry (SWASV). The sensor performance for th eon-site monitoring of Hg2+ including calibration curves, potential interference, repeatability, recovery, and limit of detection (LOD) was then fully evaluated in a landfill leachate matrix. This study was the first to investigate direct electrochemical Hg2+ detection in a real wastewater sample.

2. Materials and Methods

2.1. AuNP-Chitosan Electrode Sensor Fabrication

AuNP-chitosan composite film was electrochemically deposited on a carbon screen-printed electrode (SPE) (RRPE1001C, Pine Research Instrumentation, Durham, NC, USA) as a working electrode. The nanocomposite solution contained chloroauric acid (HAuCl4) and chitosan, based on the optimization ratio of chitosan to sensor performance used in a previous study [34]. The schematic illustration of the electrodeposition of the composite material is presented in Figure S1. The AuNP and biopolymer nanocomposite was mixed at a 1:1 ratio (12 mg chitosan in 10 mL DI solution and 0.01 M HAuCl4 solution), which was stirred for 24 h at 60 °C to achieve complete mixing. Then, 3 µL of the solution was dropped onto the surface of the carbon SPE sensor and completely deposited by electrodeposition at 6.4 mA/cm2 DC for 1 h. The electrodeposition was repeated three times to achieve sufficient deposition of AuNPs. The fabricated sensor was cleaned with DI water, dried at room temperature, and stored under ambient conditions before use.

2.2. Structure Characterization of AuNP-Biopolymer Composite

The surface structures of the fabricated SPE sensors were obtained using an Ultra 55 scanning electron microscope (SEM) (ZEISS, Oberkochen, Germany). Elemental mapping of the surface was achieved through energy-dispersive X-ray spectroscopy (EDS) using a Noran System 7 EDS with a silicon drift X-ray detector (Thermo Scientific™, Waltham, MA, USA). To characterize the surface chemical and oxidation states, X-ray photoelectron spectroscopy (XPS) was performed by a Thermo Scientific ESCALAB Xi+ X-ray Photoelectron Spectrometer Microprobe (Thermo Scientific™, Waltham, MA, USA) with a twin-crystal, micro-focusing monochromator and an Al anode. The XPS analysis chamber during measurement was held at a pressure below 1.0 × 10−7 torr. Electrochemical impedance spectroscopy (EIS) spectra were measured using a PalmSens 3 EIS potentiostat (PalmSens Compact Electrochemical Interfaces, Houten, The Netherlands) under the three-electrode system in 10 mM potassium ferricyanide (K3Fe(CN)6) solution at the frequency range 5 Hz to 50 kHz.

2.3. Chemicals and Sample Preparation

Analytical grade sodium acetate was obtained from Sigma-Aldrich (St. Louis, MO, USA) and used without further purification. To characterize the developed sensor for Hg detection, a 0.1 M acetate buffer solution (AcB) at pH 3.0 was used. We used a stock Hg solution (1 mg/mL in nitric acid, 7487-94-7, Sigma-Aldrich, St. Louis, MO, USA) ranging from 2 to 20 ppb in the electrolyte (i.e., acetate buffer). For the sensor evaluation in real wastewater, landfill leachate samples were obtained from a local landfill (Orange county landfill, Orlando, FL, USA). The characterization of the landfill leachate samples is shown in Table S1. Test water samples were prepared by spiking precalculated amounts of Hg2+ ranging from 10 to 100 ppb into the landfill leachate and the sensor performance of mercury detection was evaluated for various mercury concentrations. To adjust pH to 3, 0.1 M HCl was added to the landfill leachate sample. The samples were kept in room conditions of 23 °C with a relative humidity of 45%. The prepared Hg2+ test solutions were validated using a single quadrupole ICP-MS (iCAP-RQ, Thermo Scientific, Waltham, MA, USA).

2.4. Sensor Characterization and Electrochemical Mercury Detection

A 20 mL electrochemical cell (Compact Voltammetry Cell-Starter Kit, Pine Research Instrumentation, Durham, NC, USA) with an effective volume of 10 mL was used for Hg2+ detection, and each series was measured by a three-electrode system consisting of an AuNP-chitosan working electrode, carbon counter electrode (RRPE1001C, Pine Research Instrumentation, Durham, NC, USA), and Ag/AgCl reference electrode (MI-401, Microelectrodes, Inc., Bedford, NH, USA). We used a PalmSens 3 EIS (PalmSens Compact Electrochemical Interfaces, Houten, The Netherlands) as a potentiostat for all tests (Figure 1). Mercury detection using the developed AuNP-biopolymer SPE sensor was characterized using SWASV and CV. First, CV was performed at a scan rate of 50 mV/s in 0.1 M AcB (pH 3.0) with Hg2+ (50 ppb) to determine the potential window of the AuNP-biopolymer SPE sensor for Hg2+ detection and the potential where the Hg2+ stripping peak appears. For the Hg2+ measurement using SWASV, Hg2+ was electrochemically deposited on the working electrode at −0.8 V at pH 3.0 of 0.1 M AcB for 100 s, and the reduced Hg2+ was then stripped according to predetermined parameters (4 mV step potential, 25 mV amplitude, and 20 Hz frequency). For sequential measurements, the working electrode was cleaned at +0.8 V for 60 s to remove remnants of Hg2+ ion before the next measurement. All tests were conducted in triplicate with a value of mean ± standard deviation (SD). The LOD was calculated based on three times the signal-to-noise (S/N ratio of 3) using the Equation (1): CL = kSB/b [35], where CL signifies the detection limit, SB represents the standard deviation of blank signals, k is a parameter with a value of 3, and b is the value of the calibration curve slope.

3. Results and Discussion

3.1. Characterization and Modification of AuNP-Biopolymer Nanostructure

The surface analysis by SEM showed that the electrode surface was covered with gold nano-urchins (AuNPs) as well as some small AuNPs, as seen in Figure 2. The AuNPs were on the order of 1 μm in diameter, with many spines on the surface of each urchin-like particle. The AuNPs shown on the electrode surface were in the tens of nanometers in diameter. Elemental analysis with energy-dispersive X-ray spectroscopy (EDS) mapping confirmed that the location of the Au signal was concentrated in these structures on the electrode surface. The EDS spectra (Figure S2) showed prominent characteristic X-ray peaks from C Kα and Au M emissions. In addition to the increased surface area, the Au urchin-like structures were expected to have improved performance through several mechanisms. For example, a greater portion of this exposed surface is likely composed of high-index facets. These have been shown to increase catalytic performance [36,37] and with Miller indices (hkl) (where at least one value is greater than 1) have a greater number of low-coordinated atoms on the surface [38]. These facets have higher surface energy, and as a result, the deposition takes place preferentially on these facets during particle growth. Co-electrodeposition of Au and chitosan successfully generated these urchin-like structures in addition to the free AuNPs on the surface of the SPE sensor.
The surface chemical state was determined by XPS, as shown in Figure 3. Deconvolution of the C 1s and N 1s core-level peaks confirmed the surface adhesion of the biopolymer. The C–C peak at 284.7 eV was the dominant peak in the sample, with a signal contribution from both the chitosan polymer and the exposed regions of the carbon electrode. Deconvolution analysis shows a C–N peak at 286.4 eV and a C–O peak at 287.2 eV, which are consistent with the structure of chitosan. The peak at 289.1 eV, consistent with C=O moiety, suggests some acetylated portions of the chitosan chain or residual carbon monoxide gas. Due to the low signal/noise ratio in the N 1s region, the deconvolution of the N 1s band is less certain; deconvolution may indicate an amide moiety peak at 400.0 eV, with −NH2 and −NH3 peaks at 399.3 eV and 400.7 eV, respectively, confirming a deacetylated chitosan biopolymer on the surface. Au 4f peaks were located at 84.5 eV (4f7/2) and 88.3 eV (4f5/2), with full width at half maxima (FWHM) of 0.80 eV and 0.76, respectively. The peak and FWHM are consistent with metallic Au [39].

3.2. AuNP-Biopolymer-Coated Carbon SPE Sensor Characterization

The electron transfer properties of the developed AuNP-biopolymer-coated carbon SPE sensor were studied by electrochemical impedance spectra (EIS) (Figure 4a). A carbon SPE sensor coated only with chitosan, one coated only with AuNPs, a bare carbon SPE sensor, and a bare gold SPE sensor were also prepared and compared for reference purposes. The charge transfer resistance for the chitosan-coated carbon SPE sensor (398 Ω) was higher compared to that of the bare carbon SPE sensor (228 Ω) and the bare gold SPE sensor (70 Ω), indicating that the chitosan only-coated carbon SPE sensor had hindered charge transfer from the redox probe of [Fe(CN)6]3−/4− to the surface of the fabricated electrode. In contrast, the Nyquist plots of the AuNP-coated SPE sensors (i.e., AuNP-coated carbon SPE sensor and AuNP-biopolymer-coated carbon SPE sensor) displayed an almost straight line, showing a very fast charge transport process due to direct electron transfer. This implied that the presence of AuNP nanocomposites enabled the enhancement of electron transfer kinetics suitable for achieving a superior sensor response. The value of the charge transfer resistance also depended on the dielectric properties at the electrode/electrolyte interface [40].
To investigate sensor characteristics for the oxidation and reduction of Hg2+ by varying potentials, CV was performed in 0.1 M AcB with a 50 ppb Hg2+ concentration (Figure 4b). The corresponding redox peaks current was gradually increased in the following order: chitosan-coated carbon SPE sensor, AuNP coated carbon SPE sensor, and AuNP-chitosan-coated carbon SPE sensor. The highest redox peak current was obtained with the AuNP-chitosan-coated carbon SPE sensor, which was attributed to the improvement in electron transport due to the metallic Au nanoparticles in chitosan. Lower redox peaks were observed for the bare electrode series (carbon and gold) due to the hindrance of electron transport because of the absence of the Au NP/chitosan composite on the working electrode. The observed findings were consistent with the results obtained from the EIS measurements. These results gave immediate evidence that the modification of the sensing interface successfully detected Hg2+.

3.3. Optimization of SWASV Parameters for Hg2+ Detection

To achieve high performance in Hg2+ detection, the electrochemical parameters for SWASV (deposition potential, deposition time, amplitude, and frequency) were optimized at a fixed Hg2+ concentration of 20 ppb in 0.1 M AcB at pH 3.0. First, the effect of the deposition potential on Hg2+-stripping peak currents was investigated between −1.0 and −0.2 V with a deposition time of 100 s, the amplitude of 25 mV and a frequency of 20 Hz [34] (Figure 5a). The current of 4.0 ± 0.1 μA at −0.6 V of deposition potential was 1.2–1.7 fold higher than other deposition potentials between −0.2 V and −1.0 V. The resulting SWASV peaks showed the maximum stripping peak current at −0.6 V, indicating the greatest reduction of Hg2+. At deposition potentials higher than −0.6 V, the current obtained from stripping Hg2+ decreased as the deposition potential increased, probably due to the inefficient deposition of Hg2+. On the other hand, at reduction potentials below −0.6 V, the reduction of Hg2+ was less efficient because the reaction began to compete with H2 generation, which typically occurs below this potential [41,42]. As such, it could be seen that the stripping peak current decreased at a deposition potential lower than −0.6 V based on the CV tests. A similar trend of deposition potential was also obtained by Rahman et al. (2019) [43]. Thus, a deposition potential of −0.6 V was chosen as an optimum potential for Hg2+.
Deposition time is known to affect the amount of Hg2+ ions [43] deposited on the AuNP-chitosan electrode, thus influencing the LOD and the overall time needed for the SWASV technique. The peak current of anodic Hg2+ stripping increased from 10.4 ± 0.6 to 39.7 ± 1.3 μA with a prolonged deposition time at the optimized deposition potential of −0.6 V (Figure 5b). The peak current for Hg2+ initially increased with deposition time, then slightly increased after 200 s probably due to Hg2+ saturation on the electrode surface. Although the sensitivity can be improved with a longer deposition time, surface saturation at high metal ion concentrations can also reduce the upper detection limit [44]. Correspondingly, the deposition time of 200 s was chosen as optimal for the fast detection of Hg2+ with high selectivity.
The effects of amplitude and frequency on th current response to Hg2+ were also investigated (Figure 5c,d). Optimal conditions for amplitude were observed at 25 mV with the highest peaks at 4.3 ± 0.04 μA. Increasing the amplitude range from 0.05 to 0.1 V resulted in higher ambient noise with a larger SD (Figure S3c) due to vibrations [45]. For frequency optimization, the peak heights increased from 20 to 100 Hz, while a higher frequency showed larger ambient noise after 80 Hz (Figure S3d) because of the frequency properties of vibration [45]. Therefore, 60 Hz was chosen as the optimal frequency for Hg2+ detection using the AuNP-chitosan SPE sensor. Overall, −0.6 V of deposition potential, 200 s of deposition time, 25 mV amplitude, and 60 Hz of frequency were selected as optimal SWASV analytical parameters for further evaluation of the sensor performance on Hg2+ detection.

3.4. Sensitivity Analysis and Lifetime

Figure 6a shows the SWASV response of AuNP-biopolymer-coated carbon SPE sensor at various Hg2+ concentrations between 0–20 ppb in 0.1M AcB (pH 3.0) under optimized parameters. The well-defined sharp anodic stripping peaks for detecting Hg2+ were located at +0.58 V, and the peak currents were linearly increased with Hg2+ concentrations. The applied optimal parameter of the AuNP-biopolymer-coated carbon SPE sensor exhibited improved sensor sensitivity (17× more sensitive than an unoptimized one) with a higher R2 value (R2 = 0.9914), indicating enhanced sensitivity and stability for Hg2+ detection (Figure S4a,b). The correlation was Ip = 0.477x + 0.0009, where Ip is the stripping peak current (µA) and x is Hg2+ concentration (ppb) (Figure 6b). The sensitivity of the sensor was ~0.1 µA/nM (0.477 µA/µg) according to the slope of the linear curve. The LOD of the AuNP-biopolymer-coated carbon SPE sensor toward Hg2+ was estimated to be 0.9 ppb according to Equation (1), which is well below the USEPA-defined limit for drinking water (2 ppb) [3]. The relative standard deviation (RSD) value of 2.27% was evaluated by replicate measurements (n = 20) of 20 ppb Hg2+ solution (Figure 6c). The linear range was also 25× larger than that of other SPE sensors (bare gold and bare carbon) (Figure S4d,f). A comparison between the developed AuNP-biopolymer-coated carbon SPE sensor and other electrochemical Hg-detecting sensors, based on a gold electrode or AuNPs previously reported in the literature, is listed in Table 1. The AuNP-biopolymer-coated carbon SPE sensor had a lower Hg2+ detection range compared to those of previous studies. This is attributable to the AuNP surface, where free functional groups of amino acids (i.e., chitosan) exhibits a strong affinity for Hg2+ between Au and Hg [46]. The sensor also showed relatively lower RSD and detection limits.

3.5. Selectivity Analysis: Interference of Other Heavy Metals in Hg2+ Detection

Metal ions commonly found in water containing Hg2+ are known to compromise the accuracy of electrochemical measurements [48]. Thus, the influence of competing metal ions was investigated in two-fold concentrations over the analyte (i.e., 40 ppb) interfering species (Zn2+, Cd2+, Pb2+, and Cu2+) in 20 ppb Hg2+ in 0.1 M AcB (pH 3.0). Figure 7 shows a similar range of sensor responses in the presence of these heavy-metal ions. Most of these (Zn2+, Cd2+, and Pb2+) did not interfere with the detection of Hg2+. A slight decrease in the current value (~8.7%) was shown in the presence of Cu2+, a major interfering species because of its close potential to Hg2+ measurement [52]. To further validate this, we performed an experiment by mixing ions with a fixed concentration of Hg2+ at 20 ppb. No decrease in peak current or interfering voltammetric peaks was observed in the solution. The developed AuNP-biopolymer-coated carbon SPE sensor was then applied to the Hg2+ detection in landfill leachate as an example of a practical analytical application.

3.6. Applications for Landfill Leachate Environment

To evaluate sensor behavior in a real wastewater matrix, landfill leachate samples were collected on site (Orange County Landfill, Orlando, FL, USA) without pretreatment except for filtration through 0.45 μm Polyethersulfone (PES) filters. The pH of the landfill leachate water was ~7.7 and included high heavy-metal and ionic concentration (Table S1). The pH was adjusted to 3.0 using 0.1 M HCl to have predominant species of Hg2+ in the range of pH 1–3 [53]. The original landfill leachate sample was directly used for the baseline curve and electrolyte due to a high NaCl concentration (13,800 µS/cm). Application of the AuNP-biopolymer-coated carbon SPE sensor in this experiment resulted in well-defined peaks at +0.6V for a range of Hg2+ concentrations, 0–100 ppb, for the landfill leachate, under which, the anodic peak potential shifted towards the positive due to the positive electrocatalytic activity of the Hg2+ ions with the sensor. This phenomenon is common as adsorption products tend to shift the peak potential positively [54]. Hg2+ sensitivity and the LOD of AuNP-biopolymer-coated carbon SPE sensor showed 0.089 μA/ppb (R2 = 0.993) and 1.69 ppb, respectively (Figure 8a,b). The reproducibility experiments conducted using landfill leachate spiked with Hg2+ (50 ppb) achieved 15 stable, successive measurements (Figure 8c) with an RSD of 5.1%. The concentrations of Hg2+ in the spiked landfill leachate sample were validated using ICP-MS and compared with the results of our method (Table 2). The recoveries were in the range of 98–108%, indicating an acceptable performance for practical applications.

4. Conclusions

We demonstrated direct Hg2+ detection in landfill leachate using a newly developed AuNP-biopolymer carbon SPE sensor. The combination of the good conductivity of AuNP-chitosan and its strong adhesion to Hg2+ gave the sensor high sensitivity and selectivity for Hg2+ determination with a detection limit of 0.9 ppb in 0.1 M AcB. The sensor was successfully applied to determine Hg2+ in real landfill leachate samples directly with recovery ranging from 98 to 108%. The LOD and RSD for 15 consecutive measurements of the fabricated sensor in the landfill leachate samples were 1.69 ppb and 5.1%, respectively. The minimal interference in the presence of other heavy metal ions was observed in the detection of Hg2+ ions using the sensor. Overall, the developed AuNP-biopolymer carbon SPE sensor is expected to demonstrate reliable Hg2+ sensing in wastewater including landfill leachate.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/mi12060649/s1: Figure S1: Schematic of the AuNP-biopolymer-coated carbon SPE fabrication process; Figure S2: XPS and SEM-EDS spectra; Figure S3: Effect of deposition potential, deposition time, amplitude, and frequency on the anodic stripping peak current of Hg2+ using an AuNP-biopolymer-coated carbon SPE; Figure S4: SWASV and calibration curves of Hg2+ using different electrodes; Table S1: Heavy-metal ion concentrations in mining wastewater and soil leachate samples are shown in Table S1; Characterization of landfill leachate.

Author Contributions

J.-H.H. and W.H.L. conceived and designed the experiments. J.-H.H., D.F. and J.S. performed the experiments and analyzed the data. J.-H.H., D.F., J.S., V.A., L.Z. and W.H.L. wrote the manuscript. J.-H.H. and W.H.L. supervised the project. All authors discussed the results and contributed to the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Department of Energy (DOE) (Grant Number: DE-AC09-08SR22470) and the Korea Agency for Infrastructure Technology Advancement (KAIA) grant funded by the Ministry of Land, Infrastructure, and Transport (Grant Number: 21UGCP-B157945-02).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Nolan, E.M.; Lippard, S.J. Tools and tactics for the optical detection of mercuric ion. Chem. Rev. 2008, 108, 3443–3480. [Google Scholar] [CrossRef]
  2. Chu, P.; Porcella, D. Mercury stack emissions from US electric utility power plants. Water Air Soil Pollut. 1995, 80, 135–144. [Google Scholar] [CrossRef]
  3. US EPA. Mercury Update: Impact on Fish Advisories; EPA-823-F-01-011; US EPA: Washington, DC, USA, 2001; pp. 1–10. [Google Scholar]
  4. Gworek, B.; Bemowska-Kałabun, O.; Kijeńska, M.; Wrzosek-Jakubowska, J. Mercury in marine and oceanic waters—A review. Water Air Soil Pollut. 2016, 227, 1–19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Björkman, L.; Lundekvam, B.F.; Lægreid, T.; Bertelsen, B.I.; Morild, I.; Lilleng, P.; Lind, B.; Palm, B.; Vahter, M. Mercury in human brain, blood, muscle and toenails in relation to exposure: An autopsy study. Environ. Health 2007, 6, 1–14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Ullrich, S.M.; Tanton, T.W.; Abdrashitova, S.A. Mercury in the aquatic environment: A review of factors affecting methylation. Crit. Rev. Environ. Sci. Technol. 2001, 31, 241–293. [Google Scholar] [CrossRef]
  7. Tchounwou, P.B.; Yedjou, C.G.; Patlolla, A.K.; Sutton, D.J. Heavy metal toxicity and the environment. Mol. Clin. Environ. Toxicol. 2012, 101, 133–164. [Google Scholar]
  8. Hoang, C.V.; Oyama, M.; Saito, O.; Aono, M.; Nagao, T. Monitoring the presence of ionic mercury in environmental water by plasmon-enhanced infrared spectroscopy. Sci. Rep. 2013, 3, 1–6. [Google Scholar] [CrossRef] [PubMed]
  9. Davidson, P.W.; Myers, G.J.; Weiss, B. Mercury exposure and child development outcomes. Pediatrics 2004, 113, 1023–1029. [Google Scholar]
  10. Weinstein, M.; Bernstein, S. Pink ladies: Mercury poisoning in twin girls. Cmaj 2003, 168, 201. [Google Scholar]
  11. Genchi, G.; Sinicropi, M.S.; Carocci, A.; Lauria, G.; Catalano, A. Mercury exposure and heart diseases. Int. J. Environ. Res. Public Health 2017, 14, 74. [Google Scholar] [CrossRef] [Green Version]
  12. Fernandes Azevedo, B.; Barros Furieri, L.; Peçanha, F.M.; Wiggers, G.A.; Frizera Vassallo, P.; Ronacher Simões, M.; Fiorim, J.; Rossi de Batista, P.; Fioresi, M.; Rossoni, L. Toxic effects of mercury on the cardiovascular and central nervous systems. J. Biomed. Biotechnol. Bioeng. 2012, 2012. [Google Scholar] [CrossRef] [PubMed]
  13. Alves, G.M.; Magalhães, J.M.; Soares, H.M. Voltammetric quantification of Zn and Cu, together with Hg and Pb, based on a gold microwire electrode, in a wider spectrum of surface waters. Electroanalysis 2013, 25, 493–502. [Google Scholar] [CrossRef]
  14. Sanchez-Rodas, D.; Corns, W.; Chen, B.; Stockwell, P. Atomic fluorescence spectrometry: A suitable detection technique in speciation studies for arsenic, selenium, antimony and mercury. J. Anal. At. Spectrom. 2010, 25, 933–946. [Google Scholar] [CrossRef] [Green Version]
  15. Passariello, B.; Barbaro, M.; Quaresima, S.; Casciello, A.; Marabini, A. Determination of mercury by inductively coupled plasma—Mass spectrometry. Microchem. J. 1996, 54, 348–354. [Google Scholar] [CrossRef]
  16. Margui, E.; Kregsamer, P.; Hidalgo, M.; Tapias, J.; Queralt, I.; Streli, C. Analytical approaches for Hg determination in wastewater samples by means of total reflection X-ray fluorescence spectrometry. Talanta 2010, 82, 821–827. [Google Scholar] [CrossRef]
  17. Leopold, K.; Harwardt, L.; Schuster, M.; Schlemmer, G. A new fully automated on-line digestion system for ultra trace analysis of mercury in natural waters by means of FI-CV-AFS. Talanta 2008, 76, 382–388. [Google Scholar] [CrossRef]
  18. Pyhtilä, H.; Perämäki, P.; Piispanen, J.; Niemelä, M.; Suoranta, T.; Starr, M.; Nieminen, T.; Kantola, M.; Ukonmaanaho, L. Development and optimization of a method for detecting low mercury concentrations in humic-rich natural water samples using a CV-ICP-MS technique. Microchem. J. 2012, 103, 165–169. [Google Scholar] [CrossRef]
  19. Wei, Y.; Yang, R.; Yu, X.-Y.; Wang, L.; Liu, J.-H.; Huang, X.-J. Stripping voltammetry study of ultra-trace toxic metal ions on highly selectively adsorptive porous magnesium oxide nanoflowers. Analyst 2012, 137, 2183–2191. [Google Scholar] [CrossRef] [PubMed]
  20. Zaib, M.; Athar, M.M.; Saeed, A.; Farooq, U. Electrochemical determination of inorganic mercury and arsenic—A review. Biosens. Bioelectron. 2015, 74, 895–908. [Google Scholar] [CrossRef]
  21. Huber, J.; Leopold, K. Nanomaterial-based strategies for enhanced mercury trace analysis in environmental and drinking waters. TrAC Trends Anal. Chem. 2016, 80, 280–292. [Google Scholar] [CrossRef]
  22. Gao, C.; Huang, X.-J. Voltammetric determination of mercury (II). TrAC Trends Anal. Chem. 2013, 51, 1–12. [Google Scholar] [CrossRef]
  23. Tung, T.T.; Nine, M.J.; Krebsz, M.; Pasinszki, T.; Coghlan, C.J.; Tran, D.N.; Losic, D. Recent advances in sensing applications of graphene assemblies and their composites. Adv. Funct. Mater. 2017, 27, 1702891. [Google Scholar] [CrossRef]
  24. Lin, Y.; Peng, Y.; Di, J. Electrochemical detection of Hg (II) ions based on nanoporous gold nanoparticles modified indium tin oxide electrode. Sens. Actuators B Chem. 2015, 220, 1086–1090. [Google Scholar] [CrossRef]
  25. Fu, X.-C.; Wu, J.; Nie, L.; Xie, C.-G.; Liu, J.-H.; Huang, X.-J. Electropolymerized surface ion imprinting films on a gold nanoparticles/single-wall carbon nanotube nanohybrids modified glassy carbon electrode for electrochemical detection of trace mercury (II) in water. Anal. Chim. Acta 2012, 720, 29–37. [Google Scholar] [CrossRef] [PubMed]
  26. Lu, X.; Dong, X.; Zhang, K.; Zhang, Y. An ultrasensitive electrochemical mercury (II) ion biosensor based on a glassy carbon electrode modified with multi-walled carbon nanotubes and gold nanoparticles. Anal. Methods 2012, 4, 3326–3331. [Google Scholar] [CrossRef]
  27. Somé, I.T.; Sakira, A.K.; Mertens, D.; Ronkart, S.N.; Kauffmann, J.-M. Determination of groundwater mercury (II) content using a disposable gold modified screen printed carbon electrode. Talanta 2016, 152, 335–340. [Google Scholar] [CrossRef] [PubMed]
  28. Takale, B.S.; Bao, M.; Yamamoto, Y. Gold nanoparticle (AuNPs) and gold nanopore (AuNPore) catalysts in organic synthesis. Org. Biomol. Chem. 2014, 12, 2005–2027. [Google Scholar] [CrossRef]
  29. Hezard, T.; Fajerwerg, K.; Evrard, D.; Collière, V.; Behra, P.; Gros, P. Gold nanoparticles electrodeposited on glassy carbon using cyclic voltammetry: Application to Hg (II) trace analysis. J. Electroanal. Chem. 2012, 664, 46–52. [Google Scholar] [CrossRef] [Green Version]
  30. Zanini, V.I.P.; Gimenez, R.E.; Pérez, O.E.L.; de Mishima, B.A.L.; Borsarelli, C.D. Enhancement of amperometric response to tryptophan by proton relay effect of chitosan adsorbed on glassy carbon electrode. Sens. Actuators B Chem. 2015, 209, 391–398. [Google Scholar] [CrossRef]
  31. Luo, X.; Zeng, J.; Liu, S.; Zhang, L. An effective and recyclable adsorbent for the removal of heavy metal ions from aqueous system: Magnetic chitosan/cellulose microspheres. Bioresour. Technol. 2015, 194, 403–406. [Google Scholar] [CrossRef]
  32. Ghalkhani, M.; Shahrokhian, S. Adsorptive stripping differential pulse voltammetric determination of mebendazole at a graphene nanosheets and carbon nanospheres/chitosan modified glassy carbon electrode. Sens. Actuators B Chem. 2013, 185, 669–674. [Google Scholar] [CrossRef]
  33. Amanulla, B.; Subbu, H.K.R.; Ramaraj, S.K. A sonochemical synthesis of cyclodextrin functionalized Au-FeNPs for colorimetric detection of Cr6+ in different industrial waste water. Ultrason. Sonochem. 2018, 42, 747–753. [Google Scholar] [CrossRef]
  34. Hwang, J.-H.; Wang, X.; Pathak, P.; Rex, M.M.; Cho, H.J.; Lee, W.H. Enhanced electrochemical detection of multiheavy metal ions using a biopolymer-coated planar carbon electrode. IEEE Trans. Instrum. Meas. 2019, 68, 2387–2393. [Google Scholar] [CrossRef]
  35. Lee, W.H.; Wahman, D.G.; Pressman, J.G. Amperometric carbon fiber nitrite microsensor for in situ biofilm monitoring. Sens. Actuators B Chem. 2013, 188, 1263–1269. [Google Scholar] [CrossRef]
  36. Quan, Z.; Wang, Y.; Fang, J. High-Index Faceted Noble Metal Nanocrystals. Acc. Chem. Res. 2013, 46, 191–202. [Google Scholar] [CrossRef] [PubMed]
  37. Yuanfeng, P.; Ruiyi, L.; Xiulan, S.; Guangli, W.; Zaijun, L. Highly sensitive electrochemical detection of circulating tumor DNA in human blood based on urchin-like gold nanocrystal-multiple graphene aerogel and target DNA-induced recycling double amplification strategy. Anal. Chim. Acta 2020, 1121, 17–25. [Google Scholar] [CrossRef] [PubMed]
  38. Van Santen, R.A. Complementary Structure Sensitive and Insensitive Catalytic Relationships. Acc. Chem. Res. 2009, 42, 57–66. [Google Scholar] [CrossRef] [PubMed]
  39. Alegria, E.C.B.A.; Ribeiro, A.P.C.; Mendes, M.; Ferraria, A.M.; Do Rego, A.M.B.; Pombeiro, A.J.L. Effect of Phenolic Compounds on the Synthesis of Gold Nanoparticles and its Catalytic Activity in the Reduction of Nitro Compounds. Nanomaterials 2018, 8, 320. [Google Scholar] [CrossRef] [Green Version]
  40. Tang, L.; Feng, H.; Cheng, J.; Li, J. Uniform and rich-wrinkled electrophoretic deposited graphene film: A robust electrochemical platform for TNT sensing. Chem. Commun. 2010, 46, 5882–5884. [Google Scholar] [CrossRef] [PubMed]
  41. Gong, J.; Zhou, T.; Song, D.; Zhang, L. Monodispersed Au nanoparticles decorated graphene as an enhanced sensing platform for ultrasensitive stripping voltammetric detection of mercury (II). Sens. Actuators B Chem. 2010, 150, 491–497. [Google Scholar] [CrossRef]
  42. Kang, W.; Pei, X.; Rusinek, C.A.; Bange, A.; Haynes, E.N.; Heineman, W.R.; Papautsky, I. Determination of lead with a copper-based electrochemical sensor. Anal. Chem. 2017, 89, 3345–3352. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Rahman, M.T.; Kabir, M.F.; Gurung, A.; Reza, K.M.; Pathak, R.; Ghimire, N.; Baride, A.; Wang, Z.; Kumar, M.; Qiao, Q. Graphene oxide–silver nanowire nanocomposites for enhanced sensing of Hg2+. ACS Appl. Nano Mater. 2019, 2, 4842–4851. [Google Scholar] [CrossRef]
  44. Wei, Y.; Yang, R.; Liu, J.-H.; Huang, X.-J. Selective detection toward Hg (II) and Pb (II) using polypyrrole/carbonaceous nanospheres modified screen-printed electrode. Electrochim. Acta 2013, 105, 218–223. [Google Scholar] [CrossRef]
  45. Koper, N.; Leston, L.; Baker, T.M.; Curry, C.; Rosa, P. Effects of ambient noise on detectability and localization of avian songs and tones by observers in grasslands. Ecol. Evol. 2016, 6, 245–255. [Google Scholar] [CrossRef] [Green Version]
  46. Rex, M.; Hernandez, F.E.; Campiglia, A.D. Pushing the limits of mercury sensors with gold nanorods. Anal. Chem. 2006, 78, 445–451. [Google Scholar] [CrossRef]
  47. Li, D.; Li, J.; Jia, X.; Wang, E. Gold nanoparticles decorated carbon fiber mat as a novel sensing platform for sensitive detection of Hg (II). Electrochem. Commun. 2014, 42, 30–33. [Google Scholar] [CrossRef]
  48. Matlou, G.G.; Nkosi, D.; Pillay, K.; Arotiba, O. Electrochemical detection of Hg (II) in water using self-assembled single walled carbon nanotube-poly (m-amino benzene sulfonic acid) on gold electrode. Sens. Biosens. Res. 2016, 10, 27–33. [Google Scholar] [CrossRef] [Green Version]
  49. Wei, J.; Yang, D.; Chen, H.; Gao, Y.; Li, H. Stripping voltammetric determination of mercury (II) based on SWCNT-PhSH modified gold electrode. Sens. Actuators B Chem. 2014, 190, 968–974. [Google Scholar] [CrossRef]
  50. Bernalte, E.; Sánchez, C.M.; Gil, E.P. Determination of mercury in ambient water samples by anodic stripping voltammetry on screen-printed gold electrodes. Anal. Chim. Acta 2011, 689, 60–64. [Google Scholar] [CrossRef]
  51. Safavi, A.; Farjami, E. Construction of a carbon nanocomposite electrode based on amino acids functionalized gold nanoparticles for trace electrochemical detection of mercury. Anal. Chim. Acta 2011, 688, 43–48. [Google Scholar] [CrossRef]
  52. Punrat, E.; Chuanuwatanakul, S.; Kaneta, T.; Motomizu, S.; Chailapakul, O. Method development for the determination of mercury (II) by sequential injection/anodic stripping voltammetry using an in situ gold-film screen-printed carbon electrode. J. Electroanal. Chem. 2014, 727, 78–83. [Google Scholar] [CrossRef]
  53. Alguacil, F.J.; López, F.A. Adsorption Processing for the Removal of Toxic Hg (II) from Liquid Effluents: Advances in the 2019 Year. Metals 2020, 10, 412. [Google Scholar] [CrossRef] [Green Version]
  54. Gao, W.; Nyein, H.Y.; Shahpar, Z.; Fahad, H.M.; Chen, K.; Emaminejad, S.; Gao, Y.; Tai, L.-C.; Ota, H.; Wu, E. Wearable microsensor array for multiplexed heavy metal monitoring of body fluids. Acs Sens. 2016, 1, 866–874. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Measurement setup schematic for testing an AuNP-biopolymer-coated carbon SPE.
Figure 1. Measurement setup schematic for testing an AuNP-biopolymer-coated carbon SPE.
Micromachines 12 00649 g001
Figure 2. SEM and EDS mapping of AuNP-biopolymer composite film electrodeposited on a carbon electrode. The electron map obtained with the secondary detector is shown (a), with the corresponding carbon intensity map (b), and gold intensity map (c). All scale bars are 1 μm.
Figure 2. SEM and EDS mapping of AuNP-biopolymer composite film electrodeposited on a carbon electrode. The electron map obtained with the secondary detector is shown (a), with the corresponding carbon intensity map (b), and gold intensity map (c). All scale bars are 1 μm.
Micromachines 12 00649 g002
Figure 3. XPS of electrodeposited AuNP-biopolymer nanocomposite. (a) Au 4f spectra, (b) C 1s spectra, and (c) N 1s spectra.
Figure 3. XPS of electrodeposited AuNP-biopolymer nanocomposite. (a) Au 4f spectra, (b) C 1s spectra, and (c) N 1s spectra.
Micromachines 12 00649 g003
Figure 4. (a) Nyquist diagrams and (b) CV responses of a bare carbon SPE sensor, a bare gold SPE sensor, a biopolymer modified carbon SPE sensor, and an AuNP-biopolymer nanocomposite SPE sensor.
Figure 4. (a) Nyquist diagrams and (b) CV responses of a bare carbon SPE sensor, a bare gold SPE sensor, a biopolymer modified carbon SPE sensor, and an AuNP-biopolymer nanocomposite SPE sensor.
Micromachines 12 00649 g004
Figure 5. Optimization of (a) deposition potential, (b) deposition time, (c) amplitude and (d) frequency on the stripping peak currents of Hg2+ using an AuNP-biopolymer-coated SPE sensor. Hg2+ concentration was 10 ppb (0.1M AcB at pH 3.0). Individual data points represent the standard deviation (±SD) of duplicate experiments.
Figure 5. Optimization of (a) deposition potential, (b) deposition time, (c) amplitude and (d) frequency on the stripping peak currents of Hg2+ using an AuNP-biopolymer-coated SPE sensor. Hg2+ concentration was 10 ppb (0.1M AcB at pH 3.0). Individual data points represent the standard deviation (±SD) of duplicate experiments.
Micromachines 12 00649 g005
Figure 6. (a) SWASV replies for Hg2+ determination, (b) plot of the stripping peak current vs Hg2+ concentration, and (c) reproducibility. Deposition time is 200 s with a −0.6 V deposition potential, 0.004 V potential step, 0.025 V amplitude, and 60 Hz frequency. Hg2+ concentration was 20 ppb (0.1M acetate buffer at pH 3.0). The error bars represent the standard deviation for the mean of the three replicate tests.
Figure 6. (a) SWASV replies for Hg2+ determination, (b) plot of the stripping peak current vs Hg2+ concentration, and (c) reproducibility. Deposition time is 200 s with a −0.6 V deposition potential, 0.004 V potential step, 0.025 V amplitude, and 60 Hz frequency. Hg2+ concentration was 20 ppb (0.1M acetate buffer at pH 3.0). The error bars represent the standard deviation for the mean of the three replicate tests.
Micromachines 12 00649 g006
Figure 7. Sensor response changes at a fixed Hg2+ concentration (20 ppb) in the presence of other metal ions (40 ppb) in 0.1 M AcB (pH 3.0). Mix indicates the solution containing Zn2+, Cd2+, Pb2+, and Cu2+ (40 ppb) along with Hg2+ (20 ppb). The error bars represent the standard deviation for the mean of the three replicate tests.
Figure 7. Sensor response changes at a fixed Hg2+ concentration (20 ppb) in the presence of other metal ions (40 ppb) in 0.1 M AcB (pH 3.0). Mix indicates the solution containing Zn2+, Cd2+, Pb2+, and Cu2+ (40 ppb) along with Hg2+ (20 ppb). The error bars represent the standard deviation for the mean of the three replicate tests.
Micromachines 12 00649 g007
Figure 8. Hg2+ detection using a AuNP-biopolymer-coated carbon SPE sensor in a real landfill leachate environment (a) SWASV, (b) calibration curves, and (c) reproducibility (Hg2+ concentration is 50 ppb in landfill leachate (pH 3.0)). The error bars represent the standard deviation for the mean of the three replicate tests.
Figure 8. Hg2+ detection using a AuNP-biopolymer-coated carbon SPE sensor in a real landfill leachate environment (a) SWASV, (b) calibration curves, and (c) reproducibility (Hg2+ concentration is 50 ppb in landfill leachate (pH 3.0)). The error bars represent the standard deviation for the mean of the three replicate tests.
Micromachines 12 00649 g008
Table 1. Comparison of different types of sensor performance for electrochemical Hg (II) detection.
Table 1. Comparison of different types of sensor performance for electrochemical Hg (II) detection.
ElectrodeMethodLinear RangeLOD (nM)RSD (%)Reproducibility (n)Sample Condition (Buffer Solution)Reference
AuNPs/CFME (1)DPASV a1–250 µM0.53.40.1 M HCl/pH 1[47]
np-AuNPs/ITO (2)DPASV0.5–50 nM0.152.370.1 M HCl/pH 1[24]
Au-DMAET-(SWCNT-PABS) (3)SWASV b20–250 µM63.42.7100.1 M HCl/pH 3[48]
SWCNT-PhSH/Au (4)SWASV5–90 nM3.03.870.1 M HCl/pH 1[49]
SPGE (5)SWASV5–30 µM5.54.3400.1 M HCl/pH 2[50]
AuNPs-GC (6)CV c0.64–4 µM0.420.01 M HCl/pH 2[29]
Cys-AuNPs-CILE (7)SWASV10–20,000 nM2.32.650.1 M phosphate/pH 7[51]
AuNP-biopolymer-coated carbon SPE sensorSWASV10–100 nM4.52.3200.1 M acetate/pH 3This study
a DPASV (differential pulse anodic stripping voltammetry); b SWASV (square wave anodic stripping voltammetry); c CV (cyclic voltammetry); (1) Gold Nanoparticles (AuNPs)/three-dimensional fibril-like carbon-fiber mat electrode (CFME); (2) Nanoporous gold nanoparticles (np-AuNPs)/Indium tin oxide (ITO); (3) Au-dimethyl amino ethanethiol (DMAET)-Single-walled carbon nanotube-poly (m-amino benzene sulfonic acid) (SWCNT-PABS); (4) Single-walled carbon nanotube (SWCNTs) with thiophenol/Gold (Au); (5) Screen-printed gold electrodes (SPGE); (6) Gold nanoparticles–modified glassy carbon (AuNPs-GC); (7) l-cysteine (Cys)-Au nanoparticle-Carbon ionic liquid electrode (CILE).
Table 2. Comparison of Hg2+ detection methods between AuNP-biopolymer-coated carbon SPE sensor and ICP-MS (n = 3).
Table 2. Comparison of Hg2+ detection methods between AuNP-biopolymer-coated carbon SPE sensor and ICP-MS (n = 3).
SampleHg2+ ConcentrationRecovery (%)
Add (ppb)Detection (ppb)
AuNP-Biopolymer-Coated Carbon SPE SensorICP-MS
Landfill leachate 11514.7 ± 1.815.2 ± 0.998
Landfill leachate 22021.6 ± 3.420.6 ± 1.3107.8
Landfill leachate 33031.5 ± 2.130.8 ± 1.4105
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Hwang, J.-H.; Fox, D.; Stanberry, J.; Anagnostopoulos, V.; Zhai, L.; Lee, W.H. Direct Mercury Detection in Landfill Leachate Using a Novel AuNP-Biopolymer Carbon Screen-Printed Electrode Sensor. Micromachines 2021, 12, 649. https://doi.org/10.3390/mi12060649

AMA Style

Hwang J-H, Fox D, Stanberry J, Anagnostopoulos V, Zhai L, Lee WH. Direct Mercury Detection in Landfill Leachate Using a Novel AuNP-Biopolymer Carbon Screen-Printed Electrode Sensor. Micromachines. 2021; 12(6):649. https://doi.org/10.3390/mi12060649

Chicago/Turabian Style

Hwang, Jae-Hoon, David Fox, Jordan Stanberry, Vasileios Anagnostopoulos, Lei Zhai, and Woo Hyoung Lee. 2021. "Direct Mercury Detection in Landfill Leachate Using a Novel AuNP-Biopolymer Carbon Screen-Printed Electrode Sensor" Micromachines 12, no. 6: 649. https://doi.org/10.3390/mi12060649

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