Binder-Free Modification of a Glassy Carbon Electrode by Using

May 23, 2019 - (35,36) In contrast, these AM compounds are suspected to cause potential diseases such as carcinogens, skin, and eye irritants. Herein,...
0 downloads 0 Views 7MB Size
This is an open access article published under an ACS AuthorChoice License, which permits copying and redistribution of the article or any adaptations for non-commercial purposes.

Article Cite This: ACS Omega 2019, 4, 8907−8918

http://pubs.acs.org/journal/acsodf

Binder-Free Modification of a Glassy Carbon Electrode by Using Porous Carbon for Voltammetric Determination of Nitro Isomers Shaktivel Manavalan,† Pitchaimani Veerakumar,*,‡,§ Shen-Ming Chen,*,† Keerthi Murugan,† and King-Chuen Lin*,‡,§ †

Department of Chemical Engineering and Biotechnology, National Taipei University of Technology, No. 1, Chung-Hsiao East Road, Section 3, Taipei 10608, Taiwan, ROC ‡ Department of Chemistry, National Taiwan University, No. 1, Roosevelt Road, Section 4, Taipei 10617, Taiwan, ROC § Institute of Atomic and Molecular Sciences, Academia Sinica, No. 1, Roosevelt Road, Section 4, Taipei 10617, Taiwan, ROC

Downloaded by 109.94.175.118 at 04:01:30:053 on May 26, 2019 from https://pubs.acs.org/doi/10.1021/acsomega.9b00622.

S Supporting Information *

ABSTRACT: In this study, Liquidambar formosana tree leaves have been used as a renewable biomass precursor for preparing porous carbons (PCs). The PCs were produced by pyrolysis of natural waste of leaves after 10% KOH activation under a nitrogen atmosphere and characterized by a variety of state-of-the-art techniques. The PCs possess a large surface area, micro-/mesoporosity, and functional groups on its surface. A glassy carbon electrode modified with high PCs was explored as an efficient binder-free electrocatalyst material for the voltammetric determination of nitro isomers such as 3-nitroaniline (3-NA) and 4-nitroaniline (4-NA). Under optimal experimental conditions, the electrochemical detection of 3-NA and 4-NA was found to have a wide linear range of 0.2−115.6 and 0.5−120 μM and a low detection limit of 0.0551 and 0.0326 μM, respectively, with appreciable selectivity. This route not only enhanced the benefit from biomass wastes but also reduced the cost of producing electrode materials for electrochemical sensors. Additionally, the sensor was successfully applied in the determination of nitro isomers even in the presence of other common electroactive interference and real samples analysis (beverage and pineapple jam solutions). Therefore, the proposed method is simple, rapid, stable, sensitive, specific, reproducible, and cost-effective and can be applicable for real sample detection.



carbon materials,21 but the treatment is costly and unfriendly to the environment.22 Recently, sustainable PC derived from biomass or earth-abundant renewable resources has attracted much attention because it is relatively effective, environmentally friendly, and simple in the synthesis approach.23−25 For instance, Liu et al. studied the nitrogen-doped carbon derived from biomass (grass), and the carbon-rich nanomaterial was exhibited as a promising fluorescent sensing platform for the label-free detection of Cu(II) ions.26 Huang et al. developed the amorphous oxygen-doped carbon nanosheet (O@CN) from tannin which was applied as the electrocatalyst for efficient N2 fixation to NH3 under ambient conditions.27 In addition, Wu et al. obtained the oxygen-doped hollow carbon microtubes (O-KFCNTs) derived from natural kapok fibers using high-temperature carbonization under argon flow. The O-KFCNTs were employed as an electrocatalyst for a metalfree N2 reduction reaction to convert N2 to NH3 with excellent selectivity and to achieve good electrochemical performances.28,29

INTRODUCTION In recent decades, different allotropic forms of carbon materials such as amorphous carbon, diamond, graphite, nanotubes, fullerene, porous carbon (PC), graphene-templated carbons, and so forth have been used extensively in catalysis, sensors, and energy-related applications.1−5 Among them, graphene and PC are chemically inert with high surface area. These physicochemical properties make them efficient to be used particularly in supercapacitors, solar cells, and electrochemical sensors.6−8 PC owns unique properties, high surface areas, and large tunable pore channels, showing advantages in fast diffusion of molecules and the rapid electron transfer process, thus having an extensive application in electrochemical sensors and so forth.9,10 Thus far, various synthetic methods such as chemical vapor deposition,11 arc discharge,12 laser ablation,13 and chemical methods14 have been developed to achieve high-quality carbon nanomaterials. More recently, the synthesis of high surface area carbon for electrochemical sensing applications are turned to natural biowaste materials.15−18 In earlier dates, the activated coke and coal are used as the nonrenewable sources to generate high surface carbon,19,20 but they do not draw much interest. Afterward, organic/ inorganic moieties are adopted as the precursors to synthesize © 2019 American Chemical Society

Received: March 5, 2019 Accepted: May 10, 2019 Published: May 23, 2019 8907

DOI: 10.1021/acsomega.9b00622 ACS Omega 2019, 4, 8907−8918

ACS Omega

Article

Liquidambar formosana is commonly known as the Chinese sweetgum or formosan gum, and its crown is pyramid-shaped around 40−60 feet wide that becomes more oval as the tree ages.30 Its leaves are star-shaped with a fragrant odor when crushed. Their colors change from green in the spring and summer to purple, red, or yellow in the fall. Thousands of tons of dead leaves are produced each year. However, it is easy to identify in the winter by the one-inch fruit balls called as Lu Lu Tong that hang from its branches, and its fruits have been used as traditional Chinese medicine.31 It grows preferentially in East Asia, and its wood is second only to oak as commercial hardwood.32 However, the tree resin was once made into gum that was used by confederate doctors to treat dysentery in the troops. Some organic substances were extracted from the sweetgum leaves to be used as medicines as well as essential oils were extracted to treat skin infections and other ailments.33 Their dead leaves are also enriched in lignocellulosic materials.34 The abundance of lignocelluloses in the sweetgum tree makes it a useful raw material for the fabrication of activated carbon (AC). Herein, we took the dead leaves of the tree as a precursor to derive PC. Aromatic amines (AMs) such as aniline and its derivatives (ortho-, para-, and meta-nitroanilines) are commonly utilized in chemical, pharmaceutical, polymer, plastic, and dye industries in which the azo (−NN−) components are useful in foods and textile applications.35,36 In contrast, these AM compounds are suspected to cause potential diseases such as carcinogens, skin, and eye irritants. Herein, we developed electrochemical sensing based upon carbon materials for fast detection of AMs in drinking and environmental waters.37−39 This electrochemical method was reported to be superior to traditional sensing and monitoring techniques such as capillary zone electrophoresis, high-performance liquid chromatography, planar electrochromatography, solid-phase extraction, and spectrophotometry methods.40−44 In this study, we aim to develop a binder-free, metal-free, high-sensitive electrochemical determination of 3-nitroaniline (3-NA) and 4-nitroaniline (4-NA) in an aqueous medium. The as-fabricated carbon materials showed high surface area, longterm stability, and easy substrate penetration structures for 3NA and 4-NA molecules. The PC-modified electrodes were demonstrated to have remarkable sensitivity with an excellent detection limit for 3-NA and 4-NA. The selectivity, stability, reproducibility, interference effect, and analysis in real samples were also examined. Moreover, such prepared PC-modified glassy carbon electrode (GCE) materials show unique properties and remarkable electrochemical sensing performances.

Figure 1. (a) XRD patterns and (b) Raman spectra of the PC samples.

and KOH-activated (400 °C and 900 °C) PC samples were compared for their crystallinity and phase and characterized by XRD patterns (Figure S1 of the Supporting Information). In Figure 1b, two typical carbon peaks at 1366 and 1606 cm−1 were identified in the Raman spectra for the synthesized samples, corresponding to D (defect, sp3, A1g symmetry) and G (graphitic, sp2, E2g vibrations) bands, respectively.46 Usually, an intensity ratio of ID/IG is used to characterize the extent of the disordered carbon presence in carbon samples,47 yielding 1.01, 0.99, 0.99, and 0.98 for PC600, PC700, PC800, and PC900 samples, respectively, indicating that the synthesized samples have a less amount of defective sites in the structure as the carbonization temperature increased.48 The resulting higher extent of the graphitic phase, associated with the conducting phase of the carbon, could facilitate electron transfer or enhance its electrocatalytic activity. N2 Adsorption−Desorption. The surface area and porosity of the PC samples were studied by nitrogen adsorption−desorption analysis. All samples displayed a typical type-IV isotherm,49 as shown in Figure 2a−d. The hysteresis loop shown in a pressure range between 0.48 and 0.99 is categorized as type-H4, which can be attributed to the capillary condensation taking place in the mesopores.50 More detailed pore parameters of PCs are shown in Table S1 (Supporting Information). When the activation temperature increased from 600 to 900 °C, the Brunauer−Emmett−Teller surface area (SBET) increased from 688.5 to 1210.4 m2 g−1, and the total pore volume (Vtot) rises from 0.82 to 0.98 cm3 g−1. While comparing these results, the KOH activation is found to be effective to increase the surface area by forming a porous structure inside the carbon framework. Figure 2a1−d1 shows the pore size distributions of as-prepared samples. Both micropores/mesopores persist in all carbon samples, consistent with the results of the N2 isotherms.51 For comparison, the yield of different samples prepared by the method of chemical activation with KOH is summarized in Table S1 (Supporting Information). The results clearly show that as activation temperature increases, the yield of O, H, S, and N decreases, but the yield of C increases. At higher temperature, the aromatic condensation may take place among the adjacent molecules, thus generating gaseous products from the carbonized char to increase the C yield. Additionally, the elemental analysis (EA) evidenced the presence of the rich C and O elements in the L. formosana leaves collected from different places (see Table S2, Supporting Information). Structural Properties. The filed emission-scanning electron microscopy (FE-SEM) micrographs of PCs are displayed in Figure 3. In Figure 3a,b, the formation of carbon particles with a large smooth surface could be seen in the PC600 sample, confirming that the carbonization process at 600 °C in



RESULTS AND DISCUSSION XRD and Raman Analysis. The X-ray diffraction (XRD) patterns of as-prepared carbons are shown in Figure 1a. All samples exhibited two broad peaks with low intensity at 2θ = 25.3° and 43.4°, which can be indexed to (002) and (100) spacings of the graphitized turbostratic carbon structure, respectively. The resulting (002) crystalline plane of graphite indicates the presence of graphitization in all samples. A minor peak at 78.3° appeared corresponding to the (110) crystalline plane of graphite following the KOH activation process in the PC600, PC700, PC800, and PC900 samples. The graphitization of PC was apparently enhanced after the KOH treatment, and the enhanced crystallinity helped improve the conductivity of the carbon material.45 The raw sample, PC (without activation), 8908

DOI: 10.1021/acsomega.9b00622 ACS Omega 2019, 4, 8907−8918

ACS Omega

Article

Figure 2. (a−d) Nitrogen adsorption−desorption isotherms and (a1−d1) the pore size distributions for prepared PCs with variable activation temperatures.

powder at above 650 °C. Finally, K2CO3 was decomposed into K2O and CO2 at higher temperatures.53 K2O was reduced by carbon to produce metallic potassium (K) at a temperature above 700 °C, which was then intercalated to the carbon matrix. After washing, this intercalated metallic K was eliminated, leading to the formation of pore structures; the treatment processes are provided in Figure S2 of the Supporting Information. Additionally, the microstructure of tree leaves-derived carbons was examined using a field emission transmission electron microscope (FE-TEM). Typical FE-TEM images of PC900 are presented in Figure 4. The PC900 sample exhibited a

Figure 3. SEM micrographs of (a,b) porous PC600, (c,d) PC700, (e,f) PC800, and (g,h) PC900 samples. Figure 4. (a−c) FE-TEM images of PC900 at different magnifications and the (d) SAED pattern.

an N2 atmosphere could convert the dried leaves to a valuable carbon product. At higher temperature, the carbon products (PC700, PC800, and PC900) showed a rough surface and highly porous sheet-like morphology with interconnected pores arranged in an irregular manner,52 which are displayed in Figure 3c−h. The formation of porous PCs was attributed to the mixed treatment of 25 mL of 10% KOH solution with the PC powders. KOH was melted at the lower temperature, and then K2CO3 was produced from the reaction of KOH and PC

cottony structure, consisting of highly interconnected 3D hollow PC layers that overlap with each other (Figure 4a−c). The selected area electron diffraction (SAED) pattern (Figure 4d) further shows two clear rings, revealing the polycrystalline nature of PC900. The two weak diffraction rings for the (200) and (101) facet reflections of graphitized carbon reveal that the 8909

DOI: 10.1021/acsomega.9b00622 ACS Omega 2019, 4, 8907−8918

ACS Omega

Article

Figure 5. (a) XPS survey spectra of PC600, PC700, PC800, and PC900, and (b−e) the corresponding spectra of O 1s.

and further, the oxygen surface groups should concentrate on the surface of ACs. In fact, XPS has been used for determination of the surface oxygen content in AC. These results are smaller than the O content obtained via EA analysis, suggesting that the O content on surface of the KOH ACs decreases with respect to the untreated carbon. Most of the surface of the ACs belongs to internal pores (micro/ mesopores), which are thus not probed by XPS. This may explain the difference in oxygen content determination between XPS and CHN analysis. When the activation temperature increases, the reaction between carbon and KOH becomes more violent in PCs, resulting in carbon oxidation such that the carbon content is reduced and that is why the oxygen content increased after the activation at 900 °C (see Table S3). FT-IR Study. The surface of carbon materials contains abundant functional groups among which the CO-type groups are predominant. These functional groups present on the surface are analyzed by Fourier transform infrared (FT-IR), showing the spectra of carbon activated at different temperatures (600−900 °C) in Figure S4, Supporting Information. The broad absorption band at 3400 cm−1 indicated the presence of both free and hydrogen-bonded −OH groups on the surface.19 The definite bands at 1628−1710 cm−1 may be assigned to CO stretching vibrations of aldehydes, ketones, lactones, or carboxyl groups.24 However, the band intensity decreased after activation, especially at higher temperatures. It can be noticed that the peak at 1222 cm−1 is observed for all the carbons, as caused by the CO group (a stretch of C−O). The association/dissociation of surface functional groups would determine the density of the surface charge for electrostatic interactions and reactive sites for chemical interactions (e.g., ligand exchange) between the carbon surface and the chemical activation during pyrolysis.46 Electrochemical Studies. The PC-modified GCE was electrocatalytically analyzed by electrochemical impedance spectroscopy (EIS) using 5 mM [Fe(CN)6]3−/4− with 0.1 M KCl as the supporting electrolyte versus Ag/AgCl. The EIS profiles of the bare GCE and PC-modified GCEs are depicted in Figure S5a, Supporting Information. In addition, the charge transfer resistances depending on various electrocatalysts were

partial graphitization of PC900 may help increase electric conductivity.50 The elemental mappings of C, N, S, and O as well as the overlay images (Figure S3a−e, of the Supporting Information) were further examined to confirm that C, N, S, and O atoms are uniformly distributed on the PC900 sample, in agreement with the X-ray photoelectron spectroscopy (XPS) results (vide infra). X-ray Photoelectron Spectroscopy. The surface states and components of the as-prepared PC samples were furthermore analyzed by XPS. Survey spectra scans (0−1200 eV) were used to quantify the surface elements, of which only oxygen and carbon were present as shown in Figure 5a. As shown in Figure 5b, the O 1s spectra of the PC600 is dominated by two components with binding energies (B.Es) at 531.8 and 533.2 eV, whereas PC700 and PC800 samples show three peaks at B.Es of 532.8, 533.6, and 535.3 eV corresponding to the carbonyl (CO), hydroxyl (C−OH), and carboxyl (OC− O), respectively, as displayed in Figure 5c,d. These ACs contain significant amounts of oxygen on their surfaces.45 The O 1s peak in the PC900 sample can be deconvoluted into four contributions (Figure 5e). As reported,45 the peak at 531.8 eV is related to the COO− group in carboxylate and the OC group, and the peak at 532.6 eV corresponds to the C−OH and CO functional groups, while the peaks at 534.0 and 538.1 eV are attributed to the oxygen single bond (OC−O) in esters and carboxylic acids, and chemisorbed oxygen or water (H−O−H) molecules. The absorbed water or oxygen can be attributed to edge-dangling bonds induced by the hightemperature chemical treatment. The XPS analysis yielded the surface oxygen contents in both mass percentage (wt %) and atomic percentage (wt %) of all samples (see Table S3 of the Supporting Information). The higher O content is consistent with the formation of carbon−oxygen surface groups. In addition, the relative percentages of oxygenated functional groups increased with increasing activation temperature with KOH treatment. The C content by the XPS analysis (in Table S3) is smaller than that by EA, whereas the O contents show the opposite trend. These facts can be readily understood because the number of atomic layers in the samples amenable to the analysis is usually larger by the EA method than by XPS, 8910

DOI: 10.1021/acsomega.9b00622 ACS Omega 2019, 4, 8907−8918

ACS Omega

Article

fabricate the working electrode.54−56 However, the use of these binders can lead to structural disintegration under chemical attacks, such that reduced electrical conductivity and increased mass transfer resistance may happen. Besides, the involvement of these binders may increase the inner resistance and introduce unnecessary weight.57,58 Herein, we facilely achieved a binder-free PC900/GCE electrode material and studied the possibility of its utilization in the voltammetric determination of nitro isomers. As such, the electrochemical performances in 50 μM of nitro isomers (3-NA and 4-NA) toward PC900 is directly applied without the binder and with the binder (Nafion) GCE in N2-saturated 0.05 M phosphate buffer solution (PBS) (pH 7.0) aqueous electrolyte (Figure S6, Supporting Information). Furthermore, the CV measurements were performed to study the electrocatalytic activity of asprepared electrodes toward nitroaniline isomers such as 3-NA and 4-NA. Initially, the PC900-modified GCE sensing toward the individual analytes, 100 μM of 3-NA (Figure 6a), and 100

measured. In general, semicircles at high and low frequencies may be attributed to the electron transfer and mass transferlimited processes associated with ferricyanide ions, respectively.51 While a bare GCE showed a sizable semicircle with higher charge transfer resistance (Rct), the PC600, PC700, PC800modified GCE yielded a notable decrease in the semicircle and Rct value. This indicates that the bare GCE gave rise to an inferior electrocatalytic activity. On the other hand, the PC900modified GCE, which exhibited the smallest semicircle and lowest Rct value compared to other electrodes showed excellent catalytic performance due to the rapid electron transfer process on the carbon surface layers (see Figure S5a, Supporting Information). Several works25,39,46 have been reported that higher surface areas and larger pore structures are favorable for electrocatalytic applications because the mass transport of the electrolyte is facilitated to obtain higher catalytic currents. In this work, the PC900-modified GCE possesses larger pore structures that enable the electrolyte solution to flow into/out of PC more easily, thus resulting in better electrocatalytic performance. In addition, AC contains a variety of surface functionalities, especially oxygen-based functional groups, which may undergo electrochemically redox reactions. Hydrophilic functionalities (−OH, −COOH, −CO, etc.) are likely to increase the wettability of the electrode and further facilitate the electrochemical reactions (Scheme 1). The related redox reactions on the surface of PC were proposed as follows:52,53 Scheme 1. Proposed Redox Reactions on the Surface of PC

Cyclic voltammetry (CV) is a viable electrochemical technique which is used to measure the electrochemical properties of an analyte or system. The CV profiles of KOH AC samples (PC600, PC700, PC800, and PC900) were carried out at a scan rate of 100 mV s−1 using potassium ferricyanide as a redox probe, as presented in Figure S5b, Supporting Information. Owing to its low electrical conductivity, carbon restricted the electron transfer probe ([Fe(CN)6]3−/4−) toward the surface of the GCE. Furthermore, the scan rate influence of the high surface area PC900-modified GCE was studied from 20 to 300 mV s−1 by CV (Figure S5c, Supporting Information) showing that the process is not irreversible because there is no significant shift in the position of the peak potentials at different scan rates. Notably, the total current obtained at PC900-modified GCE is larger than that at the bare GCE. It may arise from a larger surface area provided by PC to enhance the redox process probed by the electrode. However, the peak current obtained at the GCE is stronger than that at the PC900-modified GCE. The peak current as a function of the scan rate at the PC900modified GCE is shown in Figure S5d, Supporting Information, yielding a linear relationship between oxidation peak current and the square root of the scan rate at the developed electrode. Generally, most of the AC materials require the use of polymer binders, such as poly(vinylene fluoride) and Nafion to

Figure 6. CV profiles for the bare GCE and different modified electrodes (PC600, PC700, PC800, and PC900) in the presence of 100 μM of (a) 3-NA, (b) 4-NA, and (c) 3-NA and 4-NA with the N2saturated 0.05 M PBS (pH 7.0) electrolyte. Inset of the panels (a1− c1) indicate the corresponding histogram of the reduction current with different modified electrodes, and (d) CVs obtained for the PC900-modified GCE at varied pHs from 3.0 to 11.0 in the presence of 100 μM of the 3-NA and 4-NA mixture. The insets of the panels (d1 and d2) correspond to the 3-NA and 4-NA peak current (Ipc) and peak potential (Epc) at the PC900-modified GCE, respectively. All measurements were recorded at a scan rate of 50 mV s−1 under N2saturated 0.05 M PBS solution.

μM (Figure 6b) of 4-NA in N2-saturated 0.05 M PBS (pH 7.0) at a scan rate of 50 mV s−1 in the potential range of +0.4 to −1.0 V versus Ag/AgCl was examined. Then, CV measurements in the co-presence of 100 μM 3-NA and 4-NA were also performed (Figure 6c). In the case of individual detection of analytes (3-NA and 4NA), a sharp reduction peak appeared at the bare GCE and PC-modified electrodes. Notably, the bare GCE showed an irreversible reduction, in contrast to the quasi-irreversible reduction with PC600, PC700, and PC800-modified GCE. On the other hand, a distinct reduction peak potential at −0.649 to −0.664 V versus Ag/AgCl was observed with the PC900modified GCE to reduce 3-NA irreversibly to 3-hydroxylami8911

DOI: 10.1021/acsomega.9b00622 ACS Omega 2019, 4, 8907−8918

ACS Omega

Article

(μM) − 0.3148 (R2 = 0.9922) and Ipa (μA) = −0.0517 (μM) − 0.2297 (R2 = 0.9926), respectively. This indicates that a maximum oxidation peak current was observed with an electrolyte pH of 7.0 over the PC900-modified GCE. As such, pH 7.0 was chosen for the supporting electrolyte for subsequent electrochemical sensing studies. Proposed Mechanism for 3-NA and 4-NA. The electrochemical mechanism of 3-NA and 4-NA at the PCmodified GCE is displayed in Scheme 2. It is well known that the pathway of nitro compounds detection involves successive four-electron transfer and protonation steps.62,64

noaniline (3-HAA). An additional oxidation peak in the potential range 0.164−0.181 V was observed as attributed to the reversible redox of the 3-HAA to 3-nitrosobenzene (3NSB).59 The inset in panel Figure 6a1 shows the histogram of the reduction current of 3-NA with different modified electrodes. Similar conclusions may be drawn for the detection of 4-NA (100 μM) in N2-saturated 0.05 M PBS (pH 7.0) at a scan rate of 50 mV s−1 for the bare GCE and PC-modified GCEs as shown in Figure 6b. Note that the modified electrodes provided a sharp reduction peak in the potential of −0.667 V, which is attributed to the irreversible reduction of the 4-NA to 4-hydroxylaminoaniline (4-HAA).60−62 Various PC-modified electrodes and their corresponding reduction currents were examined as displayed in the inset of Figure 6b1. We found a redox peak pair detected at the potential range of +0.133 to +0.116 V, which is ascribed to the reversible redox of 4-HAA to 4-nitrosobenzene. Thus, the GCE and PC-modified GCEs exhibited profound electrocatalytic reduction current at lower peak potential, indeed serving as a superior electrocatalyst for individual detection of 3-NA and 4-NA. Then, the capabilities of various modified electrodes for simultaneous detection of 3-NA and 4-NA were also assessed. Figure 6c shows the CVs obtained for the bare GCE and PCmodified electrodes in the presence of 100 μM concentration of 3-NA and 4-NA in N2-saturated PBS (0.05 M, pH 7.0) at a scan rate of 50 mV s−1. The bare GCE exhibited poor redox behavior, indicating the inefficient electrocatalytic capability when 3-NA and 4-NA were mixed for simultaneous detection. In contrast, the PC-modified electrodes exhibited two sharp reduction peaks around at −0.628 to −0.652 V and −0.718 to −0.741 V versus the Ag/AgCl reference electrode, corresponding to the irreversible reduction of 3-NA and 4-NA. Apart from this result, the reversible redox behavior of 3-NA and 4-NA was observed around +0.066 to −0.253 V, indicating that the PC-modified electrode is a suitable oxidation catalyst for simultaneous detection of 3-NA and 4-NA, in good agreement with previous reports.63,64 The inset in panel Figure 6c1 shows the reduction current of 3-NA and 4-NA versus different modified electrodes with a trend similar to the individual detection of the sample. From the CV results, a small reduction peak current of 4-NA can be noticed at bare GCE due to the limited electrical conductivity or deficiency of active sites. The legible separate reduction peaks were observed for 3-NA and 4-NA in all the tested electrodes, of which the PC900 electrode yielded the highest reduction peak current for the two samples. The superior electrochemical activity for the PC-modified GCE may be attributed to the high surface area and porosity possessed. Especially, the metal-free PC900 with the functional moieties brings about fast electron transfer during the electrochemical process to provoke high electrocatalytic activity. Because the electrocatalytic performance of the electrode depends on the pH value of the supporting electrolyte, CV curves were recorded as a function of pH from 3.0 to 11.0 in the presence of 100 μM of the 3-NA and 4-NA analyte in N2saturated 0.05 M PBS, as shown in Figure 6d. With an increase in the pH values, the potential shifts toward the right direction. In addition, the reduction current of 3-NA and 4-NA increase from pH 3.0 to 7.0 and then decrease from pH 9.0 to 11.0 (Figure 6d1). The inset (Figure 6d2) displays a linear plot of the reduction potential (V) versus pH (3.0−11.0) for 3-NA and 4-NA. The linearity may be expressed by Ipa(μA) = −0.053

Scheme 2. Proposed Electrochemical Mechanism of (a) 3NA and (b) 4-NA at the PC-Modified Electrode

During the CV measurement, catalytic reduction of 3-NA tends to gain four electrons and four protons (4H+, 4e−) at first to form 3-HAA; this is an unstable intermediate, soon followed by releasing reversibly two electrons and two protons (2H+, 2e−) to form nitrosobenzene (3-NSB) and vice versa (Scheme 2a). As for the reduction reaction of 4-NA, the same processes associated with protons and/or electrons transfer are otherwise involved, as shown in Scheme 2b. Because the redox process depends largely on the position of the nitro group and pH of the electrolyte,62,63 the nitro group (−NO2) at the ortho position in 3-NA is expected to have more rapid electron transfer reaction than the nitro group at the para position in 4NA. The electrochemical process involves the formation of an intermediate compound (3-HAA or 4-HAA) at the 3- or 4position of the benzene ring due to the transfer of two electrons and two protons. In short, the electrochemical reduction of 3-NA and 4-NA is associated with an equal number of protons and/or electrons transfer. Feasibility for Individual and Simultaneous Detection. The electrocatalytic activity of the PC900-modified GCE toward the reduction of 3-NA and 4-NA at different concentrations was measured with CV performed under N2saturated 0.05 M PBS (pH 7.0) at a scan rate of 50 mV s−1. The CV measurements were recorded (Figure 7a) in the presence of 3-NA at a concentration of 50−250 μM. Under these working conditions, 3-NA exhibited a reduction peak around −0.636 V versus Ag/AgCl, which is ascribed to the reduction of 3-NA to 3-HAA, as mentioned above. Also, during the reduction process, the cathodic current increased at 8912

DOI: 10.1021/acsomega.9b00622 ACS Omega 2019, 4, 8907−8918

ACS Omega

Article

Ipc (μA) = −62.646 + 5.891 (mV1/2 s−1/2) (R2 = 0.9943) for 3‐NA Ipc (μA) = −95.446 + 8.503 (mV1/2 s−1/2) (R2 = 0.9919) for 4‐NA

This indicates that the reduction of 3-NA and 4-NA upon the PC900-modified GCE is a diffusion-controlled process.62 As for an irreversible electrode process, the relationship between the slope of Ipa versus scan rate (V s−1)1/2 is expressed by the Laviron equation as follows:63

i RT yz jij RTks zyz jij RTks zyz zzInjj Epc = E o + jjj zz + jj zzIn(v) k αnF { k αnF { k αnF { where α is the cathodic electron transfer coefficient, ks is the rate constant of the surface reaction, n is the number of electron transfer, and R, T, and F have their usual notation. The slope of the above equation yields αn of 0.524 and 0.556 for 3-NA and 4-NA, respectively. Generally, α is assumed to be 0.5 for irreversible electrode process64 and consequently, the values for the number of electrons (n) involved in the reduction of 3-NA and 4-NA is nearby 4.0. In addition, trace amounts of heteroatoms and surface functional groups may create the defects and accelerate the facile electron transfer at the electrode because the space between the electrode and electroactive species is reduced by the π−π stacking interaction between PC900 and an aromatic structure of 3-NA and 4-NA. Concentration Effect of 3-NA and 4-NA. Apart from its electrocatalytic properties, the PC900-modified GCE demonstrates a higher capacity for analysis of 3-NA (Figure 8a) and

Figure 7. CV profiles for the PC900-modified GCE in the presence of (a) 50−250 μM [3-NA], (b) 50−150 μM of [4-NA], (c) fixed [3NA] and [4-NA] in the range of 100−200 μM at a scan rate of 50 mV s−1, and (d) 100 μM of [3-NA] and [4-NA] at varied scan rates (20− 300 mV s−1); the insets (a1−c1) are the linear plots of the reduction peak current vs concentration of 3-NA, 4-NA, and their mixture, respectively; the inset (d1) indicates a linear plot of the reduction peak current vs square root of the scan rate for a mixture of 3-NA and 4NA. All measurements were recorded under the N2-saturated 0.05 M PBS (pH 7.0) electrolyte.

reversible redox of 3-HAA peaking around +0.0661 V, which might correspond to the 3-HAA reduction to 3-NSB. As shown in the inset (Figure 7a1), a linear plot of the reduction peak current at the potential −0.661 V versus concentration of 3-NA was obtained with a correlation coefficient (R2) better than 0.9969. This potential value was then selected for further detection experiments. Likewise, the reduction peak current as a function of 4-NA concentration (50−150 μM) was acquired with the PC900-modified GCE in N2-saturated PBS (Figure 7b). A linear plot with a correlation coefficient (R2) of 0.9971 is displayed in Figure 7b1. Obviously, the reduction peak current increases proportionally when the 3-NA and 4-NA concentration increases, suggesting that the PC900-modified GCE exhibits effective electrocatalytic ability toward the reduction of 3-NA and 4-NA, owing to high active surface area, excellent electronic feature, and good conductivity. Figure 7c shows that the CVs obtained for the Ipc increases with increasing a mixed concentration of 3-NA and 4-NA. The corresponding CV results using the PC900-modified GCE in N2-saturated 0.05 M PBS (pH 7.0) and the mixture of 3-NA and 4-NA (100−200 μM) led to linear dependences of cathodic current (Ipc) versus their concentrations, as displayed in the inset (Figure 7c1). The linearity yielded R2 values of 3NA (0.9913) and 4-NA (0.9885). These dependencies provide the basis for an amperometric method for the simultaneous determination of 3-NA and 4-NA by voltammetry. Kinetic Studies. To examine the effect of the scan rate on the reduction peak current, the electrochemical performance over the PC900-modified GCE was conducted in the presence of analytes at different scan rates (20−300 mV s−1; Figure 7d). The inset of Figure 7d1 shows that the reduction peak currents increased linearly with the square root of the scan rates. The linear regression equations are found to be

Figure 8. DPV response of the PC900-modified GCE in N2-saturated 0.05 M PBS (pH 7.0) under varied analyte concentrations of (a) [3NA] 0.2−115.6 μM and (b) [4-NA] 0.5−120 μM. The inset panels (a1 and b1) show the corresponding calibration plots of the reduction peak current vs concentration for 3-NA and 4-NA, respectively.

4-NA (Figure 8b). This electroanalytical study was conducted using differential pulse voltammetry (DPV). All electrochemical parameters were initially optimized in N2-saturated 0.05 M PBS (pH 7.0), containing 0.2−115.6 μM of 3-NA. The inset of Figure 8a shows the calibration curve for 3-NA and 4NA reduction. The cathodic current peak height of 3-NA increases linearly with its concentration in the range of 0.2− 115.6 μM of 3-NA. The minimum concentration calculated to be detected by the proposed electrode is 0.0551 μM (S/N = 3) in terms of limit of detection (LOD). Furthermore, we calculated the linear regression: Ipa (μA) = −0.099 (μM) − 1.402 with the sensitivity of 4.371 μA·μM−1 cm−2. Likewise, Figure 8a displays the DPV profiles of 4-NA, showing that the reduction current increases with increasing 4-NA concentrations from 0.5 to 120 μM. The inset in panel Figure 8b 8913

DOI: 10.1021/acsomega.9b00622 ACS Omega 2019, 4, 8907−8918

ACS Omega

Article

Figure 9. DPV profiles of the PC900-modified GCE in N2-saturated 0.05 M PBS (pH 7.0) under varied concentrations of (a) 3-NA and 50 μM 4NA, (b) varied concentrations of 4-NA and 50 μM 3-NA, and (c) equal concentrations of 3-NA and 4-NA. The inset panels (a1−c1) show the corresponding calibration plots of the reduction peak current vs concentrations.

Table 1. Comparisons of Analytical Parameters for Detection of 3-NA among Various Modified Electrodes electrode

method

electrolyte/pH

linear range (μM)

LOD (μM)

refs

poly-DHCBAQSa/Grb-NFc/GCE SWCNTsd-ILGe/GCEf PC900/GCE

DPVg DPV DPV

B-Rh/2.0 PBS/7.0 PBS/7.0

0.36−4.34 0.01−7 0.2−115.6

0.152 0.008 0.0551

62 65 this work

a f

7-[(2,4-Dihydroxy-5-carboxybenzene)azo]-8-hydroxyquinoline-5-sulfonic acid. GCEs. gDifferential pulse voltammetry. hBritton−Robinson buffer solution.

b

Graphene. cNafion.

d

Single-walled CNTs. eIonic liquid gel.

Table 2. Comparisons of Analytical Parameters for Detection of 4-NA among Various Modified Electrodes modified electrodes

method

electrolyte/pH

linear range (μM)

LOD (μM)

refs

poly-DHCBAQSa/Grb-NFc/GCEd Ag electrode GCE BMIMPF6-SWNTe/GCE SWNT/GCE DTDf/AgNPs/CPEg AgNPs-POSSh/rGOi/GCE Ag/CPE GC/TPDTj-SiO2k/AgNPs ZnO NRsl/FTOm Ag@Pd NRDs/L@ERGOn Pt−Pd NPs/CNTs−rGO Ag/CPE PC900/GCE

DPVo DPV DPV LSVp DPV DPV DPV DPV SWVq AMPr DPV DPV CV DPV

B-Rs/2.0 B-R/2.0 B-R/2.0 PBS/4.0 PBS/4.0 PBS/4.0 PBS/7.2 B-R/2.0 PBS/7.2 PBS/7.0 0.5 M KCl 0.5 M KCl B-R/2.0 PBS/7.0

0.36−4.34 0.008−1000 2−100 0.01−7 0.5−10 1.0−100 0.7−551.6 0.08−100 1−80 0.001−0.012 0.007−832.6

0.137 0.00474 0.2 0.008 0.2 0.23 0.36 0.0418 0.50 0.5 0.000018 0.004

0.5−120

0.0326

62 66 67 68 68 69 70 71 72 73 74 75 76 this work

a

7-[(2,4-Dihydroxy-5-carboxybenzene)azo]-8-hydroxyquinoline-5-sulfonic acid. bGraphene. cNafion. dGCE. e1-Butyl-3-methylimidazolium hexafluorophosphate with single-walled CNT gel. f6,7,9,10,17,18,19,20,21,22-Decahydrodibenzo[h,r][1,4,7,11,15]trioxadiazacyclonanodecine16,23-dione. gCarbon paste electrode. hPolyhedral oligomeric silsesquioxane. iReduced graphene oxide. jN-[3-(Trimethoxysilyl)propyl]diethylenetriamine. kSilica. lZinc oxide nanorods. mFluorine-doped tin oxide. nElectrochemically reduced graphene oxide. oDifferential pulse voltammetry. pLinear sweep voltammetry. qSquare wave voltammetry. rAmperometry. sBritton−Robinson buffer solution.

is Ipa (μA) = −0.070 (μM) − 1.708, and LOD was estimated to be 0.53 μM (S/N = 3). Furthermore, the reduction current of 3-NA was directly proportional to its concentrations, while the reduction current of 4-NA barely changed. Likewise, Figure 9b reveals that the reduction current of 4-NA increased with its concentration, while the concentration of 3-NA was at constant (50 μM). In addition, as shown in Figure 9b1, the linear regression equation Ipa (μA) = −0.122 (μM) − 3.123 was obtained between the reduction current and concentration of 4-NA with an LOD of 0.7356 μM (S/N = 3). The analytical sensing performance of the proposed sensor was compared with previously reported 3-NA sensors, and the results are summarized in Table 1. Furthermore, we explore the sensor for practical applications, and the simultaneous determination of 3-NA and 4-NA of the PC900-modified GCE was evaluated in the presence of equal concentration of both analytes (3-NA and 4-NA). The DPV profile shows that the individual reduction peaks sharply

shows that the linear range was obtained for reduction current as a function of concentration of 4-NA following a linear regression: Ipa (μA) = −0.129 (μM) − 1.015 with a low detection limit (LOD) of 0.0326 μM (S/N = 3) and a sensitivity of 5.792 μA·μM−1 cm−2. The good analytical performance of the sensor is due to the fast electron mobility and strong electron donating nature of the high surface area PC matrix toward the reduction of nitro isomers. Selectivity and Stability of the Sensor. Furthermore, we evaluate the selectivity of the sensor for detection of 3-NA in the presence of 50 μM 4-NA (Figure 9a) or 4-NA in the presence of 50 μM 3-NA (Figure 9b); DPV measurements were performed using the PC900-modified GCE. Obviously, 3NA increases its reduction current with concentration, while the concentration of 4-NA was fixed constant (50 μM). As shown in the Figure 9a1 inset, the reduction peak current of 3NA is linear over the concentrations ranging from 2.0 to 69 μM with an R2 value of 0.9963. The linear regression equation 8914

DOI: 10.1021/acsomega.9b00622 ACS Omega 2019, 4, 8907−8918

ACS Omega

Article

Figure 10. DPV profiles for PC900-modified GCE (a) 3-NA- and 4-NA-spiked beverage drink sample, (b) pineapple jam sample, and (c) 0.5 mM of potential interference species and 50 μM and 75 μM of 3-NA and 4-NA. All measurements were recorded under the N2-saturated 0.05 M PBS (pH 7.0) electrolyte.

system, which was stored at 4 °C for each experiment. The unchanged curve shape and current after 30 days of successive scanning indicated that the fabricated carbon electrode material (see Figure S7, Supporting Information) can be used for nitro isomer long-term sensing applications with excellent electrochemical stability. Additionally, the selectivity of the sensor was examined using a PC900-modified GCE by DPV in N2-saturated PBS (0.05 M, pH 7.0) solution with 3NA and 4-NA together in the co-presence of 0.5 mM in other possible electroactive interferences such as Cu2+, Hg2+, Pd2+, Fe2+, Cd2+, Cl−, and NO32−. Obviously, nearly no change in their respective peak potential was found with additions of 3NA and 4-NA (Figure 10c); these results further reveal that the PC900-modified electrode may be exploited as a sensitive and selective electrochemical sensor for 3-NA and 4-NA.

increases when their concentrations were increased equally, as displayed in Figure 9c. The inset in the panel of Figure 9c1 displays two linear regression equations Ipa (μA) = −0.050 (μM) − 2.057 and Ipa (μA) = −0.077 (μM) − 3.635 for 3-NA and 4-NA, respectively. However, the LOD of 1.2371 μM (S/ N = 3) for 3-NA and 1.7254 for 4-NA are calculated. PC-based electrodes have been widely used in voltammetric studies because of low cost, availability, stability, and easy modification of the carbon morphology. There are a number of carbon-based electrodes such as GCE,66 carbon nanotubes (CNTs),67 and most recently, graphene.73,74 Moreover, the proposed PC900-modified GCE-based sensor indeed exhibits superior analytical performances with a wide linear range and LOD compared to other previously reported electrode materials for electrochemical detection of 4-NA, as listed in Table 2.61,65−75 Practical Analysis, Reproducibility, and Selectivity. The practicality of the PC900-modified GCE was tested for 3NA and 4-NA contained in different real samples such as cold drinks (Figure 10a) and pineapple jam samples (Figure 10b). The experiments were carried out with the additions of 3-NA and 4-NA at the spiking level of 5, 10, and 20 μM in blank cold drinks and pineapple jam samples. As summarized in Table 3,



CONCLUSIONS In summary, activated tree leaves-derived PC has been obtained from L. formosana and applied as the binder-free electrode material for the detection of nitroaniline isomers (3NA and 4-NA). The PC samples at varied carbonization temperatures have been characterized by XRD, Raman, BET, SEM, FE-TEM, EDX, CV, and DPV techniques. We found that mesoporous PC900 owns a very high surface area, resulting in better electrocatalytic activities for the reduction of nitroaniline isomers (3-NA and 4-NA) than the other PCmodified electrodes. The developed portable electrochemical sensor (PC900-modified GCE) shows attractive analytical features such as good sensitivity, widely operational linear range, LOD, and excellent selectivity. The surface area effect and functional moieties on their surface are responsible for the higher sensitivity with the ultratrace detection limit. Additionally, the PC900-modified GCE demonstrates good reproducibility and excellent long-term stability. Furthermore, the presence of 3-NA and 4-NA in the beverage and jam was determined, yielding satisfactory results. Consequently, our proposed electrode can be used for the sensitive and accurate determination of nitroaniline isomers in food samples.

Table 3. Practical Analysis of PC900-Modified GCE for 3-NA + 4-NA-Spiked Samples sample

added (μM)

found (μM)

recovery (%)

beverage (orange)

10 20 30 10 20 30

9.98 20.03 30.02 10.01 20.05 30.01

99.0 101.5 101.0 100.5 102.5 100.5

pineapple jam

the recoveries of the analyte in cold drinks and pineapple jam samples are from 94.80 to 98.87%. These results indicate that the electrochemical method proposed in this work should be suitable for sensing 3-NA and 4-NA in various food products. The reproducibility of the PC900-modified GCE in presence of 3-NA and 4-NA was evaluated showing the relative standard deviation of 6.3% (n = 10) according to the current intensity for the concentration of 100 μM 3-NA and 4-NA. Such reproducibility is highly acceptable for an electrochemical detection of 3-NA and 4-NA. The stability was inspected by acquisition of the electrode response with the same concentration of 4-NA and 3-NA within 15 days interval. The current response was found to decrease to 93% after 30 days in the N2-saturated 0.05 M PBS (pH 7.0) electrolyte



EXPERIMENTAL SECTION Chemicals. 3-NA (96%, Sigma-Aldrich) and 4-NA (96%, Sigma-Aldrich) (KOH, Acros) were obtained commercially and used without further purification. The PBS at pH 7.0 was prepared by using 0.05 M Na2HPO4 and NaH2PO4 solutions. The pH of the solutions was adjusted with 0.5 M H2SO4 and 2.0 M NaOH. A conventional three-electrode cell system contained a modified GCE (working electrode), Ag/AgCl (in saturated KCl; reference electrode), and a platinum wire (counter electrode). All other chemicals used belonged to 8915

DOI: 10.1021/acsomega.9b00622 ACS Omega 2019, 4, 8907−8918

ACS Omega

Article

Scheme 3. Schematic Diagram for the Preparation and Application of PCs



analytical grade, and all solutions were prepared by using ultrapure water (Millipore). Preparation of Tree Leaves Carbon. Dry L. formosana (sweetgum) precursors were gathered during the fall season from the campus of National Taiwan University in Taiwan. In this work, hierarchical PCs were successfully prepared by using fallen dead tree leaves as the carbon source (Scheme 3). The pretreatment of the dry leaves can be processed by the following three steps. (i) The leaves were washed with warm distilled water to remove impurities and undesirable materials and dried at room temperature for 2 days. The resulting leaf powders turn to brown color. (ii) Then, the brown powders (2.0 g) were dispersed in 25 mL of 10% KOH solution. The mixture was sonicated for 1 h to achieve homogeneity and then dried in atmospheric air conditions. The carbonization and activation processes were carried out at a single step. The dried leaf powder was heated at 600−900 °C with a heating rate of 10 °C min−1 in a tubular furnace under the N2 environment for 2 h. The samples were left to cool to ambient conditions and removed from the furnace prior to further analysis. Notably, the brown powder was totally turned into black. The obtained carbons were washed thoroughly by deionized water and then dried in an oven overnight. For comparison, some products were rinsed by a diluted hydrochloric solution (0.5 M). The resulting sample was placed in air and dried overnight at 110 °C. PC600, PC700, PC800, and PC900 were labeled for activated PC prepared from dead tree leaves carbonized at 600, 700, 800, and 900 °C. Electrochemical Measurements. A three-electrode system as mentioned above was employed for the electrochemical performance. The electrochemical analysis depended on a computerized electrochemical analyzer (CH Instruments; CHI 1205b and CHI410a) workstation. Prior to each measurement, N2 gas was purged through the electrolyte for 30 min. EIS (IM6ex ZAHNER, Kronach, Germany) was used to characterize the electrochemical properties of the as-prepared composites. To prepare PC-modified GCE, ∼2.0 mg of PC powder was dispersed in 1.0 mL of water and sonicated for 2 h. Then, 8.0 μL of the aliquot was dropped onto GCE to dry in air, followed by gentle heating in a hot air oven for 25 min without any binder used.

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsomega.9b00622. Electrochemical measurements, KOH activation, FETEM images, FT-IR spectra, EA, EIS, and CV studies, and surface area calculations (PDF)



AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected]. Phone: +886-2-23668230 (P.V.). *E-mail: [email protected]. Phone: +886-227017147 (S.-M.C.). *E-mail: [email protected]. Phone: +886-2-33661162 (K.C.L.). ORCID

Pitchaimani Veerakumar: 0000-0002-6899-9856 Shen-Ming Chen: 0000-0002-3749-1224 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS The authors are grateful for the financial support (MOST 1062113-M-027-003-MY3 to S.M.C.; MOST-102-2113-M-002009-MY3 to K.C.L.) from the Ministry of Science and Technology (MOST), Taiwan.



REFERENCES

(1) Titirici, M.-M.; White, R. J.; Brun, N.; Budarin, V. L.; Su, D. S.; del Monte, F.; Clark, J. H.; MacLachlan, M. J. Sustainable carbon materials. Chem. Soc. Rev. 2015, 44, 250−290. (2) Gu, W.; Yushin, G. Review of nanostructured carbon materials for electrochemical capacitor applications: advantages and limitations of activated carbon, carbide-derived carbon, zeolite-templated carbon, carbon aerogels, carbon nanotubes, onion-like carbon, and graphene. Wiley Interdiscip. Rev.: Energy Environ. 2014, 3, 424−473. (3) Wang, L.; Zhang, Q.; Chen, S.; Xu, F.; Chen, S.; Jia, J.; Tan, H.; Hou, H.; Song, Y. Electrochemical sensing and biosensing platform based on biomass-derived macroporous carbon materials. Anal. Chem. 2014, 86, 1414−1421. (4) De, S.; Balu, A. M.; van der Waal, J. C.; Luque, R. Biomassderived porous carbon materials: Synthesis and catalytic applications. ChemCatChem 2015, 7, 1608−1629.

8916

DOI: 10.1021/acsomega.9b00622 ACS Omega 2019, 4, 8907−8918

ACS Omega

Article

fabric for highly sensitive electrochemical sensors. Adv. Eng. Mater. 2018, 20, 1700608. (26) Liu, S.; Tian, J.; Wang, L.; Zhang, Y.; Qin, X.; Luo, Y.; Asiri, A. M.; Al-Youbi, A. O.; Sun, X. Hydrothermal treatment of grass: A lowcost, green route to nitrogen-doped, carbon-rich, photoluminescent polymer nanodots as an effective fluorescent sensing platform for label-free detection of Cu(II) ions. Adv. Mater. 2012, 24, 2037−2041. (27) Huang, H.; Xia, L.; Cao, R.; Niu, Z.; Chen, H.; Liu, Q.; Li, T.; Shi, X.; Asiri, A. M.; Sun, X. A biomass-derived carbon-based electrocatalyst for efficient N2 fixation to NH3 under ambient conditions. Chem.Eur. J. 2019, 25, 1914−1917. (28) Wu, T.; Li, P.; Wang, H.; Zhao, R.; Zhou, Q.; Kong, W.; Liu, M.; Zhang, Y.; Sun, X.; Gong, F. Biomass-derived oxygen-doped hollow carbon microtubes for electrocatalytic N2-to-NH3 fixation under ambient conditions. Chem. Commun. 2019, 55, 2684−2687. (29) Liu, S.; Tian, J.; Wang, L.; Zhang, Y.; Qin, X.; Luo, Y.; Asiri, A. M.; Al-Youbi, A. O.; Sun, X. Hydrothermal treatment of grass: A lowcost, green route to nitrogen-doped, carbon-rich, photoluminescent polymer nanodots as an effective fluorescent sensing platform for label-free detection of Cu(II) ions. Adv. Mater. 2012, 24, 2037−2041. (30) Chen, J.; Wang, W.-H.; Liu, T.-W.; Wu, F.-H.; Zheng, H.-L. Photosynthetic and antioxidant responses of Liquidambar formosana and Schima superba seedlings to sulfuric-rich and nitric-rich simulated acid rain. Plant Physiol. Biochem. 2013, 64, 41−51. (31) Wen, C.-H.; Lin, S.-S.; Chu, F.-H. Transcriptome analysis of a subtropical deciduous tree: autumn leaf senescence gene expression profile of formosan gum. Plant Cell Physiol. 2014, 56, 163−174. (32) Lingbeck, J. M.; O’Bryan, C. A.; Martin, E. M.; Adams, J. P.; Crandall, P. G. Sweetgum: An ancient source of beneficial compounds with modern benefits. Pharmacogn. Rev. 2015, 9, 1−11. (33) Konno, C.; Oshima, Y.; Hikino, H.; Yang, L.-L.; Yen, K.-Y. Antihepatotoxic Principles of Liquidambar formosana Fruits1. Planta Med. 1988, 54, 417−419. (34) Dat, N. T.; Lee, I. S.; Cai, X. F.; Shen, G.; Kim, Y. H. Oleanane triterpenoids with inhibitory activity against NFAT transcription factor from Liquidambar formosana. Biol. Pharm. Bull. 2004, 27, 426− 428. (35) Feng, Y.; Liu, S.; Liu, G.; Yao, J. Facile and fast removal of oil through porous carbon spheres derived from the fruit of Liquidambar formosana. Chemosphere 2017, 170, 68−74. (36) Pinheiro, H. M.; Touraud, E.; Thomas, O. Aromatic amines from azo dye reduction: status review with emphasis on direct UV spectrophotometric detection in textile industry wastewaters. Dyes Pigm. 2004, 61, 121−139. (37) Benigni, R.; Giuliani, A.; Franke, R.; Gruska, A. Quantitative structure−activity relationships of mutagenic and carcinogenic aromatic amines. Chem. Rev. 2000, 100, 3697−3714. (38) Ambs, S.; Neumann, H.-G. Acute and chronic toxicity of aromatic amines studied in the isolated perfused rat liver. Toxicol. Appl. Pharmacol. 1996, 139, 186−194. (39) Brüschweiler, B. J.; Merlot, C. Azo dyes in clothing textiles can be cleaved into a series of mutagenic aromatic amines which are not regulated yet. Toxicol. Appl. Pharmacol. 2017, 88, 214−226. (40) Karthikeyan, S.; Viswanathan, K.; Boopathy, R.; Maharaja, P.; Sekaran, G. Three dimensional electro catalytic oxidation of aniline by boron doped mesoporous activated carbon. J. Ind. Eng. Chem. 2015, 21, 942−950. (41) Zhao, L.; Zhu, L.; Lee, H. K. Analysis of aromatic amines in water samples by liquid−liquid−liquid microextraction with hollow fibers and high-performance liquid chromatography. J. Chromatogr. A 2002, 963, 239−248. (42) Takeda, S.; Wakida, S.-I.; Yamane, M.; Kawahara, A.; Higashi, K. Separation of aniline derivatives by micellar electrokinetic chromatography. J. Chromatogr. A 1993, 653, 109−114. (43) Nurok, D.; Koers, J. M.; Carmichael, M. A. Role of buffer concentration and applied voltage in obtaining a good separation in planar electrochromatography. J. Chromatogr. A 2003, 983, 247−253. (44) Tong, C.; Guo, Y.; Liu, W. Simultaneous determination of five nitroaniline and dinitroaniline isomers in wastewaters by solid-phase

(5) Ke, Q.; Wang, J. Graphene-based materials for supercapacitor electrodes−A review. J. Materiomics 2016, 2, 37−54. (6) Abioye, A. M.; Ani, F. N. Recent development in the production of activated carbon electrodes from agricultural waste biomass for supercapacitors: a review. Renewable Sustainable Energy Rev. 2015, 52, 1282−1293. (7) Sahasrabudhe, A.; Kapri, S.; Bhattacharyya, S. Graphitic porous carbon derived from human hair as ‘green’counter electrode in quantum dot sensitized solar cells. Carbon 2016, 107, 395−404. (8) Benzigar, M. R.; Talapaneni, S. N.; Joseph, S.; Ramadass, K.; Singh, G.; Scaranto, J.; Ravon, U.; Al-Bahily, K.; Vinu, A. Recent advances in functionalized micro and mesoporous carbon materials: synthesis and applications. Chem. Soc. Rev. 2018, 47, 2680−2721. (9) Sodtipinta, J.; Amornsakchai, T.; Pakawatpanurut, P. Nanoporous carbon derived from agro-waste pineapple leaves for supercapacitor electrode. Adv. Nat. Sci.: Nanosci. Nanotechnol. 2017, 8, 035017. (10) Veeramani, V.; Madhu, R.; Chen, S.-M.; Lou, B.-S.; Palanisamy, J.; Vasantha, V. S. Biomass-derived functional porous carbons as novel electrode material for the practical detection of biomolecules in human serum and snail hemolymph. Sci. Rep. 2015, 5, 10141. (11) Zhou, Y.; Hong, L. The growth of porous carbon fibres through in situ vapour deposition. RSC Adv. 2013, 3, 19769−19773. (12) Dai, W.; Kim, S. J.; Seong, W.-K.; Kim, S. H.; Lee, K.-R.; Kim, H.-Y.; Moon, M.-W. Porous carbon nanoparticle networks with tunable absorbability. Sci. Rep. 2013, 3, 2524. (13) Iijima, S.; Wakabayashi, T.; Achiba, Y. Structures of carbon soot prepared by laser ablation. J. Phys. Chem. 1996, 100, 5839−5843. (14) Lee, J.; Kim, J.; Hyeon, T. Recent progress in the synthesis of porous carbon materials. Adv. Mater. 2006, 18, 2073−2094. (15) Bao, S. J.; Guo, C. X.; Li, C. M. Hydrophilic porous carbon with tailored nanostructure and its sensitive hydrogen peroxide biosensor. RSC Adv. 2012, 2, 1014−1020. (16) Sha, T.; Li, X.; Liu, J.; Sun, M.; Wang, N.; Bo, X.; Guo, Y.; Hu, Z.; Zhou, M. Biomass waste derived carbon nanoballs aggregation networks-based aerogels as electrode material for electrochemical sensing. Sens. Actuators, B 2018, 277, 195−204. (17) Li, X.; Wang, Y.; Liu, J.; Sun, M.; Bo, X.; Wang, H.-L.; Zhou, M. Amperometric ascorbic acid biosensor based on carbon nanoplatelets derived from ground cherry husks. Electrochem. Commun. 2017, 82, 139−144. (18) Li, X.; Li, H.; Liu, T.; Hei, Y.; Hassan, M.; Zhang, S.; Lin, J.; Wang, T.; Bo, X.; Wang, H.-L.; Li, H.; Zhou, M. The biomass of ground cherry husks derived carbon nanoplates for electrochemical sensing. Sens. Actuators, B 2018, 255, 3248−3256. (19) Wu, M.; Zha, Q.; Qiu, J.; Han, X.; Guo, Y.; Li, Z.; Yuan, A.; Sun, X. Preparation of porous carbons from petroleum coke by different activation methods. Fuel 2005, 84, 1992−1997. (20) Wang, L.; Wang, R.; Zhao, H.; Liu, L.; Jia, D. High rate performance porous carbon prepared from coal for supercapacitors. Mater. Lett. 2015, 149, 85−88. (21) Li, S.; Cheng, C.; Liang, H.-W.; Feng, X.; Thomas, A. 2D porous carbons prepared from layered organic−inorganic hybrids and their use as oxygen-reduction electrocatalysts. Adv. Mater. 2017, 29, 1700707. (22) Shen, J.; Hou, J.; Guo, Y.; Xue, H.; Wu, G.; Zhou, B. Microstructure control of RF and carbon aerogels prepared by sol-gel process. J. Sol-Gel Sci. Technol. 2005, 36, 131−136. (23) Sangili, A.; Veerakumar, P.; Chen, S.-M.; Rajkumar, C.; Lin, K.C. Voltammetric determination of vitamin B2 by using a highly porous carbon electrode modified with palladium-copper nanoparticles. Microchim. Acta 2019, 186, 299. (24) Jain, A.; Balasubramanian, R.; Srinivasan, M. P. Production of high surface area mesoporous activated carbons from waste biomass using hydrogen peroxide-mediated hydrothermal treatment for adsorption applications. Chem. Eng. J. 2015, 273, 622−629. (25) Jiao, Y.; Cho, S. W.; Lee, S.; Kim, S. H.; Jeon, S.-Y.; Hur, K.; Yoon, S. M.; Moon, M.-W.; Wang, A. A Hierarchically porous carbon 8917

DOI: 10.1021/acsomega.9b00622 ACS Omega 2019, 4, 8907−8918

ACS Omega

Article

extraction and high-performance liquid chromatography with ultraviolet detection. Chemosphere 2010, 81, 430−435. (45) Bhoyate, S.; Ranaweera, C. K.; Zhang, C.; Morey, T.; Hyatt, M.; Kahol, P. K.; Ghimire, M.; Mishra, S. R.; Gupta, R. K. EcoFriendly and high performance supercapacitors for elevated temperature applications using recycled tea leaves. Glob. Challenges 2017, 1, 1700063. (46) Chang, J.; Gao, Z.; Wang, X.; Wu, D.; Xu, F.; Wang, X.; Guo, Y.; Jiang, K. Activated porous carbon prepared from paulownia flower for high performance supercapacitor electrodes. Electrochim. Acta 2015, 157, 290−298. (47) Ma, H.; Liu, Z.; Wang, X.; Zhang, C.; Jiang, R. Supercapacitive performance of porous carbon materials derived from tree leaves. J. Renewable Sustainable Energy 2017, 9, 044105. (48) Ahmed, S.; Parvaz, M.; Johari, R.; Rafat, M. Studies on activated carbon derived from neem (azadirachta indica) bio-waste, and its application as supercapacitor electrode. Mater. Res. Express 2018, 5, 045601. (49) Sing, K. S. W.; Everett, D.; Haul, R.; Moscou, L.; Pierotti, L.; Rouquerol, J.; Siemieniewska, T. Reporting physisorption data for gas/solid systems with special reference to the determination of surface area and porosity (Recommendations 1984). Pure Appl. Chem. 1985, 57, 603−619. (50) Biswal, M.; Banerjee, A.; Deo, M.; Ogale, S. From dead leaves to high energy density supercapacitors. Energy Environ. Sci. 2013, 6, 1249−1259. (51) Veerakumar, P.; Chen, S.-M.; Madhu, R.; Veeramani, V.; Hung, C.-T.; Liu, S.-B. Nickel nanoparticle-decorated porous carbons for highly active catalytic reduction of organic dyes and sensitive detection of Hg(II) ions. ACS Appl. Mater. Interfaces 2015, 7, 24810−24821. (52) Ayyalusamy, S.; Mishra, S.; Suryanarayanan, V. Promising postconsumer PET-derived activated carbon electrode material for nonenzymatic electrochemical determination of carbofuran hydrolysate. Sci. Rep. 2018, 8, 13151. (53) Wang, R.; Wang, P.; Yan, X.; Lang, J.; Peng, C.; Xue, Q. Promising porous carbon derived from celtuce leaves with outstanding supercapacitance and CO2 capture performance. ACS Appl. Mater. Interfaces 2012, 4, 5800−5806. (54) Bo, X.; Zhou, M. Electrochemical sensors based on ordered mesoporous carbons. In Advanced Electrode Materials; Tiwari, A., Kuralay, F., Uzun, L., Eds.; 2016; pp 213−242. (55) Park, B.-H.; Choi, J.-H. Improvement in the capacitance of a carbon electrode prepared using water-soluble polymer binder for a capacitive deionization application. Electrochim. Acta 2010, 55, 2888− 2893. (56) Rasouli, H.; Naji, L.; Hosseini, M. G. Electrochemical and electromechanical behavior of nafion-based soft actuators with PPy/ CB/MWCNT nanocomposite electrodes. RSC Adv. 2017, 7, 3190− 3203. (57) Yin, Y.; Zhao, G.; Liu, G. A sensitive electrochemical sensor using a mesoporous carbon and Nafion composite for the simultaneous detection of cadmium and lead. Int. J. Electrochem. Sci. 2017, 12, 5378−5391. (58) Mohammad, A.; Ahmad, K.; Rajak, R.; Mobin, S. M. Binder free modification of glassy carbon electrode by employing reduced graphene oxide/ZnO composite for voltammetric determination of certain nitroaromatics. Electroanalysis 2018, 30, 274−282. (59) Xu, G.; Zheng, C.; Zhang, Q.; Huang, J.; Zhao, M.; Nie, J.; Wang, X.; Wei, F. Binder-free activated carbon/carbon nanotube paper electrodes for use in supercapacitors. Nano Res. 2011, 4, 870− 881. (60) Ibrahim, A. A.; Umar, A.; Kumar, R.; Kim, S. H.; Bumajdad, A.; Baskoutas, S. Sm2O3-doped ZnO beech fern hierarchical structures for nitroaniline chemical sensor. Ceram. Int. 2016, 42, 16505−16511. (61) Ahmad, N.; Umar, A.; Kumar, R.; Alam, M. Microwave-assisted synthesis of ZnO doped CeO2 nanoparticles as potential scaffold for highly sensitive nitroaniline chemical sensor. Ceram. Int. 2016, 42, 11562−11567.

(62) Lin, X.; Ni, Y.; Kokot, S. Voltammetric analysis with the use of a novel electro-polymerised graphene-nafion film modified glassy carbon electrode: Simultaneous analysis of noxious nitroaniline isomers. J. Hazard. Mater. 2012, 243, 232−241. (63) Laviron, E. General expression of the linear potential sweep voltammogram in the case of diffusionless electrochemical systems. J. Electroanal. Chem. Interfacial Electrochem. 1979, 101, 19−28. (64) Wu, K.; Sun, Y.; Hu, S. Development of an amperometric indole-3-acetic acid sensor based on carbon nanotubes film coated glassy carbon electrode. Sens. Actuators, B 2003, 96, 658−662. (65) Zhao, F.; Liu, L.; Xiao, F.; Li, J.; Yan, R.; Fan, S.; Zeng, B. Sensitive voltammetric response of p-nitroaniline on single-wall carbon nanotube-ionic liquid gel modified glassy Carbon electrodes. Electroanalysis 2007, 19, 1387−1393. (66) Laghrib, F.; Boumya, W.; Lahrich, S.; Farahi, A.; El Haimouti, A.; El Mhammedi, M. A. Electrochemical evaluation of catalytic effect of silver in reducing 4-nitroaniline: Analytical application. J. Electroanal. Chem. 2017, 807, 82−87. (67) Pfeifer, R.; Tamiasso-Martinhon, P.; Sousa, C.; Moreira, J. C.; do Nascimento, M. A. C.; Barek, J. Differential pulse voltammetric determination of 4-nitroaniline using a glassy carbon electrode: comparative study between cathodic and anodic quantification. Monatsh. Chem. 2016, 147, 111−118. (68) Wood, T. E.; Tan, Z.; Schmoeckel, A. K.; O’Neill, D.; Atanasoski, R. Non-precious metal oxygen reduction catalyst for PEM fuel cells based on nitroaniline precursor. J. Power Sources 2008, 178, 510−516. (69) Zavar, M. H. A.; Heydari, S.; Rounaghi, G. H.; Eshghi, H.; Azizi-Toupkanloo, H. Electrochemical behavior of para-nitroaniline at a new synthetic crown ether-silver nanoparticle modified carbon paste electrode. Anal. Methods 2012, 4, 953−958. (70) Bai, W.; Sheng, Q.; Ma, X.; Zheng, J. Synthesis of silver nanoparticles based on hydrophobic interface regulation and its application of electrochemical catalysis. ACS Sustainable Chem. Eng. 2015, 3, 1600−1609. (71) Laghrib, F.; Ajermoun, N.; Hrioua, A.; Lahrich, S.; Farahi, A.; El Haimouti, A.; Bakasse, M.; El Mhammedi, M. Investigation of voltammetric behavior of 4-nitroaniline based on electrodeposition of silver particles onto graphite electrode. Ionics 2018, 1−9. (72) Rameshkumar, P.; Ramaraj, R. Electroanalysis of nitrobenzene derivatives and nitrite ions using silver nanoparticles deposited silica spheres modified electrode. J. Electroanal. Chem. 2014, 731, 72−77. (73) Ahmad, R.; Tripathy, N.; Ahn, M.-S.; Hahn, Y.-B. Development of highly-stable binder-free chemical sensor electrodes for pnitroaniline detection. J. Colloid Interface Sci. 2017, 494, 300−306. (74) Jeena, S. E.; Gnanaprakasam, P.; Selvaraju, T. Electroreduction of nitroaromatic compounds at electrochemically reduced graphene oxide supported bimetallic Ag@Pd nanorods modified electrodes. Electroanalysis 2016, 28, 1984−1991. (75) Yuan, C.-X.; Fan, Y.-R.; Tao-Zhang, H.-X.; Guo, H.-X.; Zhang, J.-X.; Wang, Y.-L.; Shan, D.-L.; Lu, X.-Q. A new electrochemical sensor of nitro aromatic compound based on three-dimensional porous Pt−Pd nanoparticles supported by graphene−multiwalled carbon nanotube composite. Biosens. Bioelectron. 2014, 58, 85−91. (76) Laghrib, F.; Lahrich, S.; Farahi, A.; Bakasse, M.; El Mhammedi, M. A. Impregnation of silver in graphite carbon using solid reaction: Electrocatalysis and detection of 4-nitroaniline. J. Electroanal. Chem. 2018, 823, 26−31.

8918

DOI: 10.1021/acsomega.9b00622 ACS Omega 2019, 4, 8907−8918