Next Article in Journal
Protocol Optimization of Proteomic Analysis of Korean Ginseng (Panax ginseng Meyer)
Previous Article in Journal
GC Analysis of Seven Seed Oils Containing Conjugated Fatty Acids
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhanced Adsorption Capacities of Fungicides Using Peanut Shell Biochar via Successive Chemical Modification with KMnO4 and KOH

1
Department of Environmental Engineering, College of Engineering, Kangwon National University, 1 Kangwondaehak-gil, Chuncheon-si 24341, Korea
2
Department of Integrated Energy and Infra System, Kangwon National University, 1 Kangwondaehak-gil, Chuncheon-si 24341, Korea
3
Corporate Collaboration Center, Sungkyunkwan University, 2066, Seobu-ro, Jangan-gu, Suwon-si 16419, Korea
*
Author to whom correspondence should be addressed.
Submission received: 24 March 2021 / Revised: 12 April 2021 / Accepted: 13 April 2021 / Published: 15 April 2021
(This article belongs to the Section Environmental Separations)

Abstract

:
This study explored the effects of peanut shell biochar (PSB) on the adsorption capacities of fungicides with and without successive chemical modifications, using KMnO4 and KOH (PSBOX-A), in order to provide a valuable understanding of their adsorption mechanisms and behaviors. To this end, the physicochemical properties of PSB and PSBOX-A were examined by using the Brunauer–Emmett–Teller method, Fourier transform infrared spectroscopy, and scanning electron microscopy with an energy dispersive X-ray spectrometer. The effects of temperature, ionic strength, and humic acids on the adsorption of fungicides, using PSB and PSBOX-A, were estimated through batch experiments. Furthermore, adsorption kinetics, isotherms, and thermodynamics were studied. The maximum adsorption capacities of fungicides by PSBOX-A were estimated to be more notable (Qmax of carbendazim = 531.2 μmol g−1, Qmax of pyrimethanil = 467.7 μmol g−1, and Qmax of tebuconazole = 495.1 μmol g−1) than PSB (Qmax of carbendazim = 92.6 μmol g−1, Qmax of pyrimethanil = 61.7 μmol g−1, and Qmax of tebuconazole = 66.7 μmol g−1). These findings suggest that successive chemical modification using KMnO4 and KOH could potentially be used to effectively fabricate PSB to remove fungicides in water-treatment processes.

1. Introduction

Fungicides are chemicals that are used to prevent fungal infections in seeds and plants. Rapid growth in the global population has led to a rise in the demand for crop yields, which has subsequently increased the use of fungicides [1]. This increase can act as a non-point source of pollution in surface water via runoff [2]. Carbendazim (CAR; methyl N-(1H-benzimidazol-2-yl)carbamate), pyrimethanil (PYR; 4,6-dimethyl-N-phenylpyrimidin-2-amine), and tebuconazole (TEB; 1-(4-chlorophenyl)-4,4-dimethyl-3-(1,2,4-triazol-1-ylmethyl)pentan-3-ol) have been frequently detected in surface waters (CAR = 0.6–6 μg L−1, PYR = 4 ng L−1, and TEB = 0.6–200 μg L−1) [3,4,5]. Even though the aforementioned fungicides are generally regarded as having low toxicity, their high chemical stability and slow degradability pose potential risks to aquatic organisms and human health [1]. However, few studies have focused on developing efficient treatments to remove fungicide pollution from water resources [6].
Recently, several researchers have recognized that adsorption is the most efficient treatment option; moreover, its operating cost for fungicide removal in water-treatment plants is low [6]. Activated carbon (AC), a representative adsorbent, is commonly utilized in the water-treatment process because it has a large specific surface area, porous structure, and a wide range of pH reactivity [7]. Moreover, various nanomaterials, including graphene and metal nanoparticles [8] and carbon nanotubes [9], have been used to remove fungicides from water. Zhaokun et al. demonstrated the efficient removal of triazole fungicides on graphene using metal nanoparticles [8]. Chen et al. fabricated magnetic carbon nanotubes for the adsorption of fungicides [9]. Although nanomaterials are beneficial due to their high efficiency and fast adsorption, their physicochemical properties (e.g., small size, persistence, and non-biodegradation) can be detrimental to humans and the environment. Furthermore, the uses of AC and nanomaterials in real-scale water-treatment processes are limited because of their high production and operation costs [10].
In recent years, new developments in waste-reuse technology, such as biochar derived from agricultural and food residues, including peanut shells and ground coffee waste, have become a global focus for research [11,12]. Furthermore, biochar can be a great alternative to AC because of its high specific surface area, porous structure, and various surface functional groups [13]. The adsorption capacity of biochar may be influenced by biomass, carbonization conditions (temperature and retention time), and chemical modification [14]. Chemical modification methods involving oxidative and alkali agents may improve the surface functionality of biochar and enhance electrostatic repulsive interactions with organic pollutants in water [15]. Peanut is a common agricultural residue in the United States. Considerable quantities (about 30 thousand metric tons) of peanuts shell are used to feed and energy resources each year [16]. The main components of peanut shells have been reported to be fiber (60–70 wt.%), cellulose (34–45 wt.%), and lignin (27–33 wt.%). The peanut shells have a cellulose–lignin layer that considerably affects the porous structure and the pyrolysis behavior [17]. In order to make peanut shells an eco-friendly and low-cost alternative adsorbent, they need to be converted into valuable resources. However, most studies have explored the adsorption mechanism by using peanut shell biochar (PSB) of heavy metals, such as Ni(II), Cd(II), Cu(II), Pb(II), and Zn(II) [16,18]. Moreover, there is a need for a comprehensive study on the effects of successive chemical modification of PSB with potassium permanganate (KMnO4) and potassium hydroxide (KOH) on the adsorption behavior of fungicides.
This study investigated the adsorption behaviors of fungicides by PSB and successive KMnO4 and KOH modified PSB (PSBOX-A) to compare the physicochemical properties of PSB and PSBOX-A (surface morphology, porous structure, bulk elemental composition, and surface functional groups). The adsorption behavior of fungicides (i.e., CAR, PYR, and TEB) by PSB and PSBOX-A was optimized and evaluated through batch experiments. The adsorption kinetics, isotherms, and thermodynamics were used to identify the primary adsorption mechanisms of CAR, PYR, and TEB, using PSB and PSBOX-A. Furthermore, the effects of temperature, ionic strength, and humic acids (HA) on the adsorption of CAR, PYR, and TEB by PSB and PSBOX-A were examined.

2. Materials and Methods

2.1. Chemicals

All chemicals and fungicides were of analytical grade. CAR (methyl N-(1H-benzimidazol-2-yl)carbamate), PYR (4,6-dimethyl-N-phenylpyrimidin-2-amine), TEB (1-(4-chlorophenyl)-4,4-dimethyl-3-(1,2,4-triazol-1-ylmethyl)pentan-3-ol), KMnO4, KOH, sodium hydroxide (NaOH), hydrochloric acid (HCl), sodium chloride (NaCl), and HA were purchased from Sigma-Aldrich (St. Louis, MO, USA). High-performance liquid chromatography (HPLC)-grade acetonitrile (ACN) was obtained from J.T. Baker (Deventer, Netherlands), and phosphoric acid (H3PO4) was purchased from FUJIFILM Wako Pure Chemical Corporation (Osaka, Japan). The use of ACN is due to its strong solvation properties for a wide range of polar and nonpolar organic solutes and advantageous properties, including low viscosity and low toxicity [19]. A mixed stock solution containing CAR, PYR, and TEB (concentration of each fungicide = 1 mmol L−1) was prepared by using ACN and deionized (DI) water (50:50, v/v). DI water was produced by using a Nanopure Water System (electrical resistivity >18.2 MΩ cm−1; Barnstead, Lake Balboa, CA, USA), and it was used to make the samples. These stock solutions were stored at 4 °C, in the dark, prior to use. The physicochemical properties of CAR, PYR, and TEB are summarized in Table 1.

2.2. Preparation of PSB and PSBOX-A

The detailed preparation method of the PSB via successive chemical modification with KMnO4 and KOH was modified in a previous study [20].

2.2.1. PSB

Peanut shells were purchased from a local grocery store in Chuncheon-si (Gangwon-do, Korea). They were rinsed several times with DI water and dried in an oven for 24 h, at 105 °C. The PSB was pyrolyzed by using a tubular furnace (PyroTech, Namyangju-si, Korea) for 2 h, at 700 °C, under N2 atmosphere (heating rate = 10 °C min−1; the N2 flow rate = 0.25 L min−1). After the entire procedure was completed, the produced PSB was rinsed with DI water and dried in an oven at 105 °C, for 12 h.

2.2.2. PSBOX-A

To enhance the surface functionality of PSB, the PSB (5 g) produced was mixed with 0.5% KMnO4 (10 mL) by stirring at 20 ± 0.5 °C for 4 h and drying in an oven, at 60 °C, for 4 h. The modified PSB with KMnO4 was pyrolyzed by using a tubular furnace at 700 °C, for 2 h, under N2 atmosphere (heating rate = 10 °C min−1; N2 flow rate = 0.25 L min−1). The modified PSB with KMnO4 was rinsed with DI water several times and dried in an oven, at 105 °C, for 12 h. Furthermore, the modified PSB with KMnO4 was added to KOH (15 g) and pyrolyzed in a tubular furnace, at 700 °C (heating rate = 10 °C min−1), for 1 h, under N2 atmosphere (flow rate = 0.25 L min−1). The PSBOX-A produced was rinsed with DI water several times and dried in an oven, at 60 °C, for 24 h. The dried PSBOX-A was passed through a 100-mesh sieve, to maintain a uniform size, after which it was stored in a desiccator prior to analysis.

2.3. Characteristics of PSB and PSBOX-A

2.3.1. Specific Surface Area and Porosity

The specific surface areas (m2 g−1) of PSB and PSBOX-A were determined by using the Brunauer–Emmett–Teller (BET) method on the N2 adsorption–desorption isotherms at 77.3 K in the relative pressure range of 0.01–0.99 (ASAP 2020 Plus, Micromeritics, GA, USA). The total pore volume (cm3 g−1) and average pore size (nm) were calculated by using the Barrett–Joyner–Halenda (BJH) method [21].

2.3.2. Elemental Compositions

The elemental compositions (i.e., carbon (C), hydrogen (H), oxygen (O), nitrogen (N), and sulfur (S)) of PSB and PSBOX-A were measured by using a EuroEA3000 elemental analyzer (EuroVector S.p.A, Via F.lli Cuzio, Italy). Ash content was calculated by subtracting the C, H, O, N, and S contents from the total amount of PSB and PSBOX-A. The atomic molar ratios of H/C and O/C + N/C were used to reveal the aromaticity and polarity of PSB and PSBOX-A, respectively [21].

2.3.3. Scanning Electron Microscopy Analysis

The surface morphologies of PSB and PSBOX-A were identified by using an S-4800 scanning electron microscope (SEM; Hitachi, Tokyo, Japan) with an energy dispersive X-ray (EDX) spectrometer.

2.3.4. Fourier Transform Infrared Spectroscopy Analysis

Fourier transform infrared (FTIR) spectroscopy with a KBr pellet was used to qualitatively analyze the surface functional groups of PSB and PSBOX-A ranging between 400 and 4000 cm−1 (Vertex 70, Bruker, Billerica, MA, USA).

2.4. Analytical Methods

The concentrations of CAR, PYR, and TEB in the samples were measured by using HPLC coupled with a column (XDB C18; ZORBAX Eclipse®, 4.6 mm × 150 mm, inner diameter = 5 µm; Agilent, Santa Clara, CA, USA) and a UVA detector (SPD-10AVP, Shimadzu, Kyoto, Japan) at 210 nm. The mobile phase (CAN: 0.05 M H3PO4 (50:50, v/v)) was operated under isocratic conditions (flow rate = 1.0 mL min−1) for 15 min. Various sample conditions were used to evaluate the effects of temperature, ionic strength, and HA on the adsorption of CAR, PYR, and TEB when using PSB and PSBOX-A (the concentration of each fungicide = 10 µmol L−1, temperature = 15–35 °C, concentrations of NaCl = 0–0.1 M, and concentration of HA = 5 mg L−1). All adsorption experiments were conducted in triplicate, using a shaking incubator (Vision Scientific, Daejeon, Korea). Then, the sample solutions were filtered by using a nominal pore size of 0.7 μm glass fiber filter (GF/F; Whatman, Maidstone, UK) to eliminate PSB and PSBOX-A.
The amounts of CAR, PYR, and TEB adsorbed per unit mass of PSB and PSBOX-A at equilibrium, Qe (μmol g−1), and removal efficiencies (%) of CAR, PYR, and TEM by PSB and PSBOX-A were calculated by using the following equations:
Q e = ( C 0 C e ) V M ,
R e m o v a l   e f f i c i e n c y   o f   f u n g i c i d e   ( % ) = C 0 C e C 0 × 100 ,
where C0 and Ce (µmol L−1) are the initial and equilibrium concentrations of CAR, PYR, and TEB, respectively. V (L) is the sample solution volume, and M (g) is the amount of PSB and PSBOX-A.
The adsorption kinetics of CAR, PYR, and TEB were examined by using the pseudo-first-order (Equation (3)) and pseudo-second-order (Equation (4)) kinetic models (adsorbent dosage = 50 mg L−1, initial concentration of each fungicide = 10 µmol L−1, contact time = 0–24 h, agitation speed = 150 rpm, temperature = 25 °C, and pH = 7.0) [22]:
Q t = Q e ( 1 1 exp ( k 1 t ) ,
Q t = Q e ( 1 1 1 + Q e k 2 t ) ,
where Qt (µmol g−1) is the amount of adsorbed CAR, PYR, and TEB on PSB and PSBOX-A at adsorption time t (h). Moreover, k1 (1/h) and k2 (g µmol−1·h) are the rate constants of the pseudo-first-order and pseudo-second-order models, respectively.
The adsorption isotherms of CAR, PYR, and TEB were investigated at five different initial concentrations (10, 20, 30, 40, and 50 µmol L−1), under fixed conditions (adsorbent dosage = 50 mg L−1, contact time = 24 h, agitation speed = 150 rpm, temperature = 25 °C, and pH = 7.0) [23].
Langmuir   Isotherm :   Q e = Q m a x K L C e 1 + K L C e ,
Freundlich   isotherm :   Q e = K F C e 1 / n ,
where Qmax (µmol g−1) is the maximum adsorption capacity in the Langmuir isotherm model, and KL (L µmol−1) is the equilibrium constant of the linearized Langmuir isotherm model. RL = 1/(1 + KLC0), derived from KL, is used to compare the adsorption affinity of the Langmuir isotherms. KF (µmol1−1/n L1/n g−1) and n (dimensionless) are constants associated with the relative maximum adsorption capacity and adsorption intensity, respectively [24].
The thermodynamic parameters of CAR, PYR, and TEB adsorption were estimated by using Equations (7)–(9) [25].
K d = Q e C e ,
G ° = R T l n ( 1000 K d ) ,
ln ( 1000 K d ) = H ° R T + S ° R ,
where Kd (L g−1) is the partition coefficient; ∆G° in (kJ mol−1), ∆H° in (kJ mol−1), and ∆S° (J mol−1·K) are the Gibbs free energy, enthalpy, and entropy, respectively. R (J mol−1·K) and T (K) are the ideal gas constant (8.314 J mol−1·K) and the absolute temperature of the sample solutions. ∆H° and ∆S° were calculated as the slope and intercept in the linear graph of ln(Kd) and 1/T, respectively.

3. Results and Discussion

3.1. Physicochemical Properties of PSB and PSBOX-A

The physicochemical properties (e.g., bulk element and ash contents, the atomic molar ratio, specific surface area, total pore volume, average pore size, and BET isotherms) associated with the adsorption capacity of PSB and PSBOX-A are presented in Figure 1 and Table 2. The bulk element and ash contents of PSB (C = 86.6%, H = 1.6%, O = 3.7%, N = 1.7%, S = 0.1%, and ash = 6.3%) and PSBOX-A (C = 82.1%, H = 1.1%, O = 4.9%, N = 1.3%, S = 0.2%, and ash = 10.5%) have revealed considerable differences due to the successive KMnO4 and KOH modification processes. Furthermore, PSBOX-A presented lower atomic molar ratios of H/C (0.16) than that of PSB (0.23), but the sum of O/C and N/C values for PSBOX-A (0.05) were similar to those of PSB (0.05). These findings indicate that PSBOX-A has a higher aromaticity (H/C value: PSB > PSBOX-A) than PSB [26]. Additionally, PSBOX-A had a higher specific surface area and total pore volume than that of PSB, but the average pore size of PSBOX-A was smaller than that of PSB. The specific surface area (1977.6 m2 g−1) and total pore volume (0.12 cm3 g−1) of PSBOX-A were significantly greater than those of PSB (specific surface area = 93.9 m2 g−1 and total pore volume = 0.04 cm3 g−1), given that the porous structures were changed by the successive chemical modification using KMnO4 and KOH [11]. However, the average pore size of PSBOX-A (3.4 nm) was not significantly different from that of PSB (3.8 nm). Therefore, the CAR, PYR, and TEB adsorption capacities of PSBOX-A were expected to be enhanced compared with those of PSB.
The SEM images of PSB and PSBOX-A are shown in Figure 2. The surface of PSB is smooth and flat, but PSBOX-A has porous and uneven structures. These results indicate that chemical modification using KMnO4 and KOH might change the surface morphology of the PSB with respect to the adsorption of fungicides [27]. The SEM–EDX results of PSB and PSBOX-A demonstrated the effect of chemical modification with KMnO4 and KOH on the differences in the constituents of PSB and PSBOX-A. The PSBOX-A was composed of C (82.05%), O (17.62%), K (0.19%), and Mn (0.14%). However, PSB only consisted of C (85.58%) and O (14.42%). The presence of K and Mn atoms in PSBOX-A could explain why K and Mn were combined on the surface of PSB during the successive chemical modification using KMnO4 and KOH.
Figure 3 shows the FTIR spectra of PSB and PSBOX-A. The IR peak at 3420 cm−1 indicates the O-H groups of alcohols [14]. The IR peaks at 1570 and 1180 cm−1 could be attributed to the vibrations of the C=C stretching of aromatic groups and the C-O stretching of phenolic and carboxylic groups, respectively [28,29]. The intensities of the IR peaks (i.e., OH alcohol groups, C=C aromatic groups, and C-O phenolic and carboxylic groups) for PSBOX-A were significantly higher than those of PSB. These observations suggest that the differences in the adsorption capacities of PSB and PSBOX-A could influence the adsorption of fungicides because of the successive chemical modifications with KMnO4 and KOH.

3.2. Effects of Absorbent Dosages

The effects of PSB and PSBOX-A dosages on the removal efficiencies of CAR, PYR, and TEB are shown in Figure 4. The adsorption capacities of CAR, PYR, and TEB using PSB and PSBOX-A continuously decreased, and the removal efficiency of fungicides on PSB and PSBOX-A gradually increased. This is because, when the initial concentrations of CAR, PYR, and TEB are fixed, the adsorption capacity per unit of the adsorbents tends to decline as the amount of PSB and PSBOX-A increases. Similar behavior was previously observed during the removal of the fungicides (e.g., triclosan and triclocarban) using Zr-based magnetic metal–organic frameworks [30]. Based on these experiments on the removal efficiency of CAR, PYR, and TEB according to the PSB and PSBOX-A dosages, 50 mg L−1 was selected as the optimal dosage and applied to the subsequent experiments.

3.3. Adsorption Kinetics of Fungicides

Figure 5 shows the adsorption kinetics of CAR, PYR, and TEB when using PSB and PSBOX-A. The fast adsorption reaction was completed within 4 h for fungicides on the PSB and PSBOX-A surfaces. After 4 h, the removal efficiencies of CAR, PYR, and TEB were increased by PSB and PSBOX-A because the activated binding sites on PSB and PSBOX-A surfaces were saturated [12]. The removal efficiencies of CAR, PYR, and TEB by PSB (removal efficiency of CAR = 7.3 ± 0.04%, removal efficiency of PYR = 3.1 ± 0.04%, and removal efficiency of TEB = 5.8 ± 0.08%) were considerably lower than those of CAR, PYR, and TEB by PSBOX-A (removal efficiency of CAR = 95.1 ± 0.1%, removal efficiency of PYR = 89.9 ± 0.02%, and removal efficiency of TEB = 92.7 ± 0.1%). These observations were attributed to the differences in the physicochemical properties (i.e., specific surface area, total pore volume, and surface functional groups) of PSB and PSBOX-A.
Table 3 presents the adsorption kinetics for CAR, PYR, and TEB when using PSB and PSBOX-A. The pseudo-second-order model (R2 = 0.999) might better explain the adsorption of fungicides when using PSB and PSBOX-A than the pseudo-first-order model (R2 = 0.882–0.985). The equilibrium adsorption capacities of the fungicides (Qe,exp) on those of PSB (Qe,exp of the PSBOX-A = 179.7–196.9 μmol g−1) were much greater than those of PSB (Qe,exp of the PSB = 6.1–15.1 μmol g−1). Furthermore, the Qe,exp values of CAR, PYR, and TEB toward the PSB and PSBOX-A demonstrate similar trends as those of the theoretical adsorption capacities (Qe, cal) calculated by the pseudo-second-order model (Qe, cal of the PSB = 6.7–15.4 μmol g−1, Qe, cal of the PSBOX-A = 184.9–202.0 μmol g−1). These results suggested that the adsorption of CAR, PYR, and TEB on PSB and PSBOX-A might be chemisorption [31].

3.4. Adsorption Isotherms of Fungicides

The adsorption of CAR, PYR, and TEB by PSB and PSBOX-A were examined by using the Langmuir and Freundlich isotherm models (Figure 6 and Table 4). The adsorption behaviors of CAR, PYR, and TEB with respect to PSB and PSBOX-A are better explained by the Langmuir isotherm model (R2 = 0.996–0.999) than the Freundlich isotherm model (R2 = 0.764–0.995). This indicated that monolayer adsorption played a key role in removing the fungicides from the homogeneous surfaces of PSB and PSBOX-A [32]. The adsorption affinities of CAR, PYR, and TEB to PSB and PSBOX-A were evaluated by using the RL values. The RL values, namely (i) RL > 1: unfavorable, (ii) RL = 1: linear, (iii) 0 < RL < 1: favorable, and (iv) RL = 0: irreversible, of the Langmuir isotherm model were used to identify the maximum adsorption capacities of CAR, PYR, and TEB by using PSB and PSBOX-A [33]. The RL values of PSB and PSBOX-A were estimated to be favorable (RL = 0.08–0.8).

3.5. Effects of pH on Adsorption of Fungicides

Figure 7 depicts the adsorption behaviors of CAR, PYR, and TEB for PSB and PSBOX-A at three different pH values (pH = 3, 7, and 11). The removal efficiencies of CAR (11.5% → 6.9%), PYR (4.8% → 2.7%), and TEB (5.6% → 3.1%) by the PSB were negative affected by the increase of pH values. However, the removal efficiencies of CAR (97.0% → 96.6%) and PYR (91.0% → 90.6%) by the PSBOX-A were not significantly different under the different pH values. The change in pH values strongly influenced the removal efficiency of TEB (94.3% → 93.2%) by the PSBOX-A due to the lower pKa value of TEB (pKa = 2.01) compared with the CAR (pKa = 4.28) and PYR (pKa = 3.44). Moreover, the deprotonated fungicides could promote the electrostatic repulsive interaction between the fungicides and the PSB under the pH value was higher than their pKa values [12]. These observations could explain that the electrostatic interactions between fungicides and absorbents under the different pH values played a key role in the adsorptions of the CAR, PYR, and TEB using the PSB and PSBOX-A.

3.6. Effects of Temperature and Thermodynamic Studies

The effects of temperature (15–35 °C) on the adsorption of fungicides using PSB and PSBOX-A are illustrated in Figure 8. The removal efficiencies of fungicides using PSB (15 °C: CAR = 4.8 ± 0.4%, PYR = 1.0 ± 0.04%, and TEB = 2.8 ± 0.03%; 25 °C: CAR = 7.2 ± 0.1%, PYR = 2.3 ± 0.2%, and TEB = 4.6 ± 0.1%; 35 °C: CAR = 8.9 ± 0.1%, PYR = 3.3 ± 0.04%, and TEB = 5.6 ± 0.2%) and PSBOX-A (15 °C: CAR = 94.3 ± 0.01%, PYR = 88.2 ± 0.03%, and TEB = 92.8 ± 0.1%; 25 °C: CAR = 94.8 ± 0.02%, PYR = 90.0 ± 0.1%, and TEB = 93.8 ± 0.02%; 35 °C: CAR = 95.8 ± 0.1%, PYR = 91.3 ± 0.1%, and TEB = 94.7 ± 0.01%) increased with temperature. A similar result was observed for the removal of herbicides by ground coffee residue biochars [14]. Moreover, the adsorption behaviors of CAR, PYR, and TEB using PSB and PSBOX-A showed the same trends (CAR > TEB > PYR) under different temperature conditions. Table 5 shows the results of the thermodynamic parameters (∆G°, ∆H°, and ∆S°) for CAR, PYR, and TEB adsorption by PSB and PSBOX-A in relation to the absolute temperatures (288–308 K). The ∆G° < 0 values indicated that the adsorption process was spontaneous. The ∆H° > 0 values designated that the adsorption reaction of CAR, PYR, and TEB on PSB and PSBOX-A was endothermic [30]. Low ∆H° values (<40 kJ mol−1) might be attributed to the physical adsorption [34]. Furthermore, ∆S° > 0 values showed conducive to adsorption stability due to irreversible adsorption [35].

3.7. Effects of NaCl and HA on Adsorption of Fungicides

The effects of ionic strength (NaCl = 0–0.1 M) on the adsorption of CAR, PYR, and TEB by PSB and PSBOX-A are compared in Figure 9. The removal efficiencies of CAR, PYR, and TEB using PSB and PSBOX-A were progressively improved with increasing ionic strengths (PSB: removal efficiency of CAR = 6.8 ± 0.07% → 10.3 ± 0.2%, removal efficiency of PYR = 2.8 ± 0.03% → 7.3 ± 0.08%, removal efficiency of TEB = 3.5 ± 0.1% → 9.3 ± 0.05%; PSBOX-A: removal efficiency of CAR = 97.1 ± 0.02% → 98.2 ± 0.01%, removal efficiency of PYR = 90.5 ± 0.01% → 91.6 ± 0.01%, removal efficiency of TEB = 94.3 ± 0.02% → 95.6 ± 0.02%). These findings reveal that reduced solubility of fungicides with increasing concentrations of NaCl can support the adsorption capacity of PSB and PSBOX-A under different ionic strength conditions (salting-out effect) [36,37].
The effects of HA on the adsorption of CAR, PYR, and TEB for PSB and PSBOX-A are presented in Figure 10. HA, macromolecular compounds with many reactive functional groups, might scavenge the binding sites generated on PSB and PSBOX-A. Thus, HA used in the batch experiments could affect the adsorption affinity of fungicides with adsorbents [38]. The removal efficiencies of CAR, PYR, and TEB by PSB are similar to the existence of HA (removal efficiency without HA: CAR = 7.1 ± 0.07%, PYR = 2.8 ± 0.09%, and TEB = 4.3 ± 0.3%; removal efficiency with HA: CAR = 6.8 ± 0.3%, PYR = 2.4 ± 0.2%, and TEB = 2.8 ± 0.1%). The removal efficiencies of CAR, PYR, and TEB using PSBOX-A are not significantly affected by the presence of HA (removal efficiency without HA: CAR = 95.2 ± 0.02%, PYR = 89.7 ± 0.07%, and TEB = 93.7 ± 0.02%; removal efficiency with HA: CAR = 93.8 ± 0.3%, PYR = 88.9 ± 0.3%, and TEB = 93.3 ± 0.07%). These observations suggested that the adsorption behavior of CAR, PYR, and TEB by PSB and PSBOX-A were outcompeted by HA adsorption [12].

4. Conclusions

This study examined the effects of successive KMnO4 and KOH modifications on the physicochemical properties of PSB associated with the adsorption behaviors of CAR, PYR, and TEB. Compared to PSB, PSBOX-A presented a higher specific surface area (1977.6 m2 g−1) and total pore volume (0.12 cm3 g−1). The PSB and PSBOX-A were well fitted with the pseudo-second-order kinetic (R2 = 0.999), and the equilibrium adsorption capacities of CAR, PYR, and TEB on PSBOX-A (Qe,exp = 179.7–196.9 μmol g−1) were greater than those of PSB (Qe,exp = 6.1–15.1 μmol g−1). These findings indicate that chemisorption plays a crucial role in the adsorption of CAR, PYR, and TEB. Moreover, the Langmuir isotherms predominantly governed the removal of fungicides from the homogeneous surfaces of PSB and PSBOX-A (monolayer adsorption, R2 = 0.996–0.999). The removal efficiencies of CAR, PYR, and TEB using PSB and PSBOX-A gradually increased with increasing temperature and NaCl concentration. Although HA could interfere with CAR, PYR, and TEB adsorptions by PSB and PSBOX-A, the removal efficiencies of fungicides by PSBOX-A (88.9–93.8%) were higher when compared to PSB (2.4–6.8%). These results suggest that successive KMnO4 and KOH modification may be a promising option in improving the adsorption capacities of PSB for removing CAR, PYR, and TEB from a real-scale water treatment plant. Future studies need to provide the information for optimum reuse conditions to regenerate the peanut shell biochars with continuous adsorption ability.

Author Contributions

Conceptualization, Y.-G.L.; methodology, Y.-G.L.; validation, Y.-G.L.; formal analysis, S.K. and C.S.; investigation, J.S. and J.K.; data curation, Y.-G.L.; writing—original draft preparation, Y.-G.L.; writing—review and editing, C.-H.L. and K.C.; supervision, K.C.; funding acquisition, G.-Y.K. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Technology Development Program (S2979139) funded by the Ministry of SMEs and Startups (MISS, Korea).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gupta, P.K. Toxicity of fungicides. In Veterinary Toxicology, 3rd ed.; Academic Press: Cambridge, MA, USA, 2018; pp. 569–580. [Google Scholar]
  2. Wang, T.; Zhang, Z.; Zhang, H.; Zhong, X.; Liu, Y.; Liao, S.; Yue, X.; Zhou, G. Sorption of carbendazim on activated carbons derived from rape straw and its mechanism. RSC Adv. 2019, 9, 41745–41754. [Google Scholar] [CrossRef] [Green Version]
  3. Merel, S.; Benzing, S.; Gleiser, C.; di Napoli-Davis, G.; Zwiener, C. Occurrence and overlooked sources of the biocide carbendazim in wastewater and surface water. Environ. Pollut. 2018, 239, 512–521. [Google Scholar] [CrossRef]
  4. Meng, Y.; Zhong, K.; Xiao, J.; Huang, Y.; Wei, Y.; Tang, L.; Chen, S.; Wu, J.; Ma, J.; Cao, Z.; et al. Exposure to pyrimethanil induces developmental toxicity and cardiotoxicity in zebrafish. Chemosphere 2020, 255, 126889. [Google Scholar] [CrossRef] [PubMed]
  5. Li, S.; Wu, Q.; Sun, Q.; Coffin, S.; Gui, W.; Zhu, G. Parental exposure to tebuconazole causes thyroid endocrine disruption in zebrafish and developmental toxicity in offspring. Aquat. Toxicol. 2019, 211, 116–123. [Google Scholar] [CrossRef]
  6. Hgeig, A.; Novaković, M.; Mihajlović, I. Sorption of carbendazim and linuron from aqueous solutions with activated carbon produced from spent coffee grounds: Equilibrium, kinetic and thermodynamic approach. J. Environ. Sci. Health Part B 2019, 54, 226–236. [Google Scholar] [CrossRef] [PubMed]
  7. Ma, X.; Ouyang, F. Adsorption properties of biomass-based activated carbon prepared with spent coffee grounds and pomelo skin by phosphoric acid activation. Appl. Surf. Sci. 2013, 268, 566–570. [Google Scholar] [CrossRef]
  8. Wang, Z.; Zhang, J.; Hu, B.; Yu, J.; Wang, J.; Guo, X. Graphene/Fe3O4 nanocomposite for effective removal of ten triazole fungicides from water solution: Tebuconazole as an example for investigation of the adsorption mechanism by experimental and molecular docking study. J. Taiwan Inst. Chem. Eng. 2019, 95, 635–642. [Google Scholar] [CrossRef]
  9. Chen, F.; Song, Z.; Nie, J.; Yu, G.; Li, Z.; Lee, M. Ionic liquid-based carbon nanotube coated magnetic nanoparticles as adsorbent for the magnetic solid phase extraction of triazole fungicides from environmental water. RSC Adv. 2016, 6, 81877–81885. [Google Scholar] [CrossRef]
  10. Yuan, M.; Tong, S.; Zhao, S.; Jia, C.Q. Adsorption of polycyclic aromatic hydrocarbons from water using petroleum coke-derived porous carbon. J. Hazard. Mater. 2010, 181, 1115–1120. [Google Scholar] [CrossRef] [PubMed]
  11. An, Q.; Jiang, Y.-Q.; Nan, H.-Y.; Yu, Y.; Jiang, J.-N. Unraveling sorption of nickel from aqueous solution by KMnO4 and KOH-modified peanut shell biochar: Implicit mechanism. Chemosphere 2019, 214, 846–854. [Google Scholar] [CrossRef]
  12. Shin, J.; Lee, Y.-G.; Lee, S.-H.; Kim, S.; Ochir, D.; Park, Y.; Kim, J.; Chon, K. Single and competitive adsorptions of micropollutants using pristine and alkali-modified biochars from spent coffee grounds. J. Hazard. Mater. 2020, 400, 123102. [Google Scholar] [CrossRef]
  13. Suo, F.; You, X.; Ma, Y.; Li, Y. Rapid removal of triazine pesticides by P doped biochar and the adsorption mechanism. Chemosphere 2019, 235, 918–925. [Google Scholar] [CrossRef]
  14. Lee, Y.-G.; Shin, J.; Kwak, J.; Kim, S.; Son, C.; Cho, K.H.; Chon, K. Effects of NaOH Activation on Adsorptive Removal of Herbicides by Biochars Prepared from Ground Coffee Residues. Energies 2021, 14, 1297. [Google Scholar] [CrossRef]
  15. Zhou, Y.; Liu, X.; Xiang, Y.; Wang, P.; Zhang, J.; Zhang, F.; Wei, J.; Luo, L.; Lei, M.; Tang, L. Modification of biochar derived from sawdust and its application in removal of tetracycline and copper from aqueous solution: Adsorption mechanism and modelling. Bioresour. Technol. 2017, 245, 266–273. [Google Scholar] [CrossRef]
  16. Wilson, K.; Yang, H.; Seo, C.W.; Marshall, W.E. Select metal adsorption by activated carbon made from peanut shells. Bioresour. Technol. 2006, 97, 2266–2270. [Google Scholar] [CrossRef]
  17. Allen, S.J.; Gan, Q.; Matthews, R.; Johnson, P.A. Mass transfer processes in the adsorption of basic dyes by peanut hulls. Ind. Eng. Chem. Res. 2005, 44, 1942–1949. [Google Scholar] [CrossRef]
  18. Johns, M.M.; Marshall, W.E.; Toles, C.A. Agricultural by-products as granular activated carbons for adsorbing dissolved metals and organics. J. Chem. Technol. Biotechnol. 1998, 71, 131–140. [Google Scholar] [CrossRef]
  19. McConvey, I.F.; Woods, D.; Lewis, M.; Gan, Q.; Nancarrow, P. The Importance of Acetonitrile in the Pharmaceutical Industry and Opportunities for its Recovery from Waste. Org. Process Res. Dev. 2012, 16, 612–624. [Google Scholar] [CrossRef] [Green Version]
  20. Fu, Y.; Shen, Y.; Zhang, Z.; Ge, X.; Chen, M. Activated bio-chars derived from rice husk via one- and two-step KOH-catalyzed pyrolysis for phenol adsorption. Sci. Total Environ. 2019, 646, 1567–1577. [Google Scholar] [CrossRef]
  21. Liu, R.; Liu, G.; Yousaf, B.; Abbas, Q. Operating conditions-induced changes in product yield and characteristics during thermal-conversion of peanut shell to biochar in relation to economic analysis. J. Clean. Prod. 2018, 193, 479–490. [Google Scholar] [CrossRef]
  22. Bhainsa, K.C.; D’souza, S. Removal of copper ions by the filamentous fungus, Rhizopus oryzae from aqueous solution. Bioresour. Technol. 2008, 99, 3829–3835. [Google Scholar] [CrossRef]
  23. Gorgievski, M.; Božić, D.; Stanković, V.; Štrbac, N.; Šerbula, S. Kinetics, equilibrium and mechanism of Cu2+, Ni2+ and Zn2+ ions biosorption using wheat straw. Ecol. Eng. 2013, 58, 113–122. [Google Scholar] [CrossRef]
  24. Hayati, B.; Mahmoodi, N.M. Modification of activated carbon by the alkaline treatment to remove the dyes from wastewater: Mechanism, isotherm and kinetic. Desalination Water Treat. 2012, 47, 322–333. [Google Scholar] [CrossRef]
  25. Li, R.; Wang, Z.; Guo, J.; Li, Y.; Zhang, H.; Zhu, J.; Xie, X. Enhanced adsorption of ciprofloxacin by KOH modified biochar derived from potato stems and leaves. Water Sci. Technol. 2018, 77, 1127–1136. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Zhang, J.; Ma, X.; Yuan, L.; Zhou, D. Comparison of adsorption behavior studies of Cd2+ by vermicompost biochar and KMnO4-modified vermicompost biochar. J. Environ. Manag. 2020, 256, 109959. [Google Scholar] [CrossRef] [PubMed]
  27. Bian, S.; Xu, S.; Yin, Z.; Liu, S.; Li, J.; Xu, S.; Zhang, Y. An Efficient Strategy for Enhancing the Adsorption Capabilities of Biochar via Sequential KMnO4-Promoted Oxidative Pyrolysis and H2O2 Oxidation. Sustainability 2021, 13, 2641. [Google Scholar] [CrossRef]
  28. Huff, M.D.; Kumar, S.; Lee, J.W. Comparative analysis of pinewood, peanut shell, and bamboo biomass derived biochars produced via hydrothermal conversion and pyrolysis. J. Environ. Manag. 2014, 146, 303–308. [Google Scholar] [CrossRef]
  29. Acemioğlu, B. Removal of a reactive dye using NaOH-activated biochar prepared from peanut shell by pyrolysis process. Int. J. Coal Prep. Util. 2019, 1–23. [Google Scholar] [CrossRef]
  30. Ahmad, M.; Lee, S.S.; Dou, X.; Mohan, D.; Sung, J.-K.; Yang, J.E.; Ok, Y.S. Effects of pyrolysis temperature on soybean stover- and peanut shell-derived biochar properties and TCE adsorption in water. Bioresour. Technol. 2012, 118, 536–544. [Google Scholar] [CrossRef]
  31. Ma, J.; Li, S.; Wu, G.; Arabi, M.; Tan, F.; Guan, Y.; Li, J.; Chen, L. Preparation of magnetic metal-organic frameworks with high binding capacity for removal of two fungicides from aqueous environments. J. Ind. Eng. Chem. 2020, 90, 178–189. [Google Scholar] [CrossRef]
  32. Kim, E.; Jung, C.; Han, J.; Her, N.; Park, C.M.; Jang, M.; Son, A.; Yoon, Y. Sorptive removal of selected emerging contaminants using biochar in aqueous solution. J. Ind. Eng. Chem. 2016, 36, 364–371. [Google Scholar] [CrossRef]
  33. Shin, J.; Lee, Y.-G.; Kwak, J.; Kim, S.; Lee, S.-H.; Park, Y.; Lee, S.-D.; Chon, K. Adsorption of radioactive strontium by pristine and magnetic biochars derived from spent coffee grounds. J. Environ. Chem. Eng. 2021, 9, 105119. [Google Scholar] [CrossRef]
  34. Fan, S.; Tang, J.; Wang, Y.; Li, H.; Zhang, H.; Tang, J.; Wang, Z.; Li, X. Biochar prepared from co-pyrolysis of municipal sewage sludge and tea waste for the adsorption of methylene blue from aqueous solutions: Kinetics, isotherm, thermodynamic and mechanism. J. Mol. Liq. 2016, 220, 432–441. [Google Scholar] [CrossRef]
  35. Son, C.; An, W.; Lee, G.; Jeong, I.; Lee, Y.-G.; Chon, K. Adsorption Characteristics of Phosphate Ions by Pristine, CaCl2 and FeCl3-Activated Biochars Originated from Tangerine Peels. Separations 2021, 8, 32. [Google Scholar] [CrossRef]
  36. Xiao, X.; Sheng, G.D.; Qiu, Y. Improved understanding of tributyltin sorption on natural and biochar-amended sediments. Environ. Toxicol. Chem. 2011, 30, 2682–2687. [Google Scholar] [CrossRef]
  37. Xiang, Y.; Xu, Z.; Zhou, Y.; Wei, Y.; Long, X.; He, Y.; Zhi, D.; Yang, J.; Luo, L. A sustainable ferromanganese biochar adsorbent for effective levofloxacin removal from aqueous medium. Chemosphere 2019, 237, 124464. [Google Scholar] [CrossRef]
  38. Cano-Aguilera, I.; Haque, N.; Morrison, G.M.; Aguilera-Alvarado, A.F.; Gutiérrez, M.; Gardea-Torresdey, J.L.; de la Rosa, G. Use of hydride generation-atomic absorption spectrometry to determine the effects of hard ions, iron salts and humic substances on arsenic sorption to sorghum biomass. Microchem. J. 2005, 81, 57–60. [Google Scholar] [CrossRef]
Figure 1. N2 adsorption-desorption isotherm curves of the (a) peanut shell biochar (PSB) and (b) KMnO4 and KOH modified PSB (PSBOX-A).
Figure 1. N2 adsorption-desorption isotherm curves of the (a) peanut shell biochar (PSB) and (b) KMnO4 and KOH modified PSB (PSBOX-A).
Separations 08 00052 g001
Figure 2. Scanning electron microscopy (SEM) images of (a) peanut shell biochar (PSB) and (b) KMnO4 and KOH modified PSB (PSBOX-A) (magnification ×800).
Figure 2. Scanning electron microscopy (SEM) images of (a) peanut shell biochar (PSB) and (b) KMnO4 and KOH modified PSB (PSBOX-A) (magnification ×800).
Separations 08 00052 g002
Figure 3. Fourier-transform infrared (FTIR) spectra of PSB and PSBOX-A.
Figure 3. Fourier-transform infrared (FTIR) spectra of PSB and PSBOX-A.
Separations 08 00052 g003
Figure 4. Effects of adsorbent dosages on the removal efficiency of CAR, PYR, and TEB by (a) PSB and (b) PSBOX-A (agitation speed = 150 rpm; initial concentration of each fungicide = 10 μM; contact time = 24 h; temperature = 25 °C; pH = 7.0).
Figure 4. Effects of adsorbent dosages on the removal efficiency of CAR, PYR, and TEB by (a) PSB and (b) PSBOX-A (agitation speed = 150 rpm; initial concentration of each fungicide = 10 μM; contact time = 24 h; temperature = 25 °C; pH = 7.0).
Separations 08 00052 g004
Figure 5. Adsorption kinetics of the selected fungicides using (a) PSB and (b) PSBOX-A (agitation speed = 150 rpm; absorbent dosage = 50 mg L−1; initial concentration of each fungicide = 10 μM; temperature = 25 °C; pH = 7.0).
Figure 5. Adsorption kinetics of the selected fungicides using (a) PSB and (b) PSBOX-A (agitation speed = 150 rpm; absorbent dosage = 50 mg L−1; initial concentration of each fungicide = 10 μM; temperature = 25 °C; pH = 7.0).
Separations 08 00052 g005
Figure 6. Adsorption isotherms of fungicides on (a) PSB and (b) PSBOX-A (agitation speed = 150 rpm; contact time = 24 h; absorbent dosage = 50 mg L−1; initial concentration of each fungicide = 10 μM; temperature = 25 °C; pH = 7.0).
Figure 6. Adsorption isotherms of fungicides on (a) PSB and (b) PSBOX-A (agitation speed = 150 rpm; contact time = 24 h; absorbent dosage = 50 mg L−1; initial concentration of each fungicide = 10 μM; temperature = 25 °C; pH = 7.0).
Separations 08 00052 g006
Figure 7. The effects of the pH on the adsorption of fungicides by (a) the PSB and (b) PSBOX-A (agitation speed = 150 rpm; contact time = 24 h; absorbent dosage = 50 mg L−1; initial concentration of each fungicide = 10 μM; temperature = 25 °C).
Figure 7. The effects of the pH on the adsorption of fungicides by (a) the PSB and (b) PSBOX-A (agitation speed = 150 rpm; contact time = 24 h; absorbent dosage = 50 mg L−1; initial concentration of each fungicide = 10 μM; temperature = 25 °C).
Separations 08 00052 g007
Figure 8. Effects of temperature on the adsorption of fungicides by (a) the PSB and (b) PSBOX-A (agitation speed = 150 rpm; contact time = 24 h; absorbent dosage = 50 mg L−1; initial concentration of each fungicide = 10 μM; pH = 7.0).
Figure 8. Effects of temperature on the adsorption of fungicides by (a) the PSB and (b) PSBOX-A (agitation speed = 150 rpm; contact time = 24 h; absorbent dosage = 50 mg L−1; initial concentration of each fungicide = 10 μM; pH = 7.0).
Separations 08 00052 g008
Figure 9. Effects of NaCl concentration on the adsorption of fungicides by (a) PSB and (b) PSBOX-A (agitation speed = 150 rpm; contact time = 24 h; absorbent dosage = 50 mg L−1; initial concentration of each fungicide = 10 μM; temperature = 25 °C; pH = 7.0).
Figure 9. Effects of NaCl concentration on the adsorption of fungicides by (a) PSB and (b) PSBOX-A (agitation speed = 150 rpm; contact time = 24 h; absorbent dosage = 50 mg L−1; initial concentration of each fungicide = 10 μM; temperature = 25 °C; pH = 7.0).
Separations 08 00052 g009
Figure 10. Effects of humic acids (HA) on the adsorption of the fungicides by (a) PSB and (b) PSBOX-A (the concentration of HA = 5 mg L−1, agitation speed = 150 rpm; contact time = 24 h; absorbent dosage = 50 mg L−1; initial concentration of each fungicide = 10 μM; temperature = 25 °C; pH = 7.0).
Figure 10. Effects of humic acids (HA) on the adsorption of the fungicides by (a) PSB and (b) PSBOX-A (the concentration of HA = 5 mg L−1, agitation speed = 150 rpm; contact time = 24 h; absorbent dosage = 50 mg L−1; initial concentration of each fungicide = 10 μM; temperature = 25 °C; pH = 7.0).
Separations 08 00052 g010
Table 1. Physicochemical properties of the selected fungicides.
Table 1. Physicochemical properties of the selected fungicides.
Compounds
(Abbreviation)
FormulaStructureMolecular Weight
(g mol−1)
Charge aLog D apKa aSolubility
in Water b
(g L−1, pH 7)
pH
3.0
pH
7.0
pH
11.0
pH
3.0
pH
7.0
pH
11.0
Carbendazim
(CAR)
C9H9N3O2 Separations 08 00052 i001191.1910−10.731.801.014.280.008
Pyrimethanil
(PYR)
C12H13N3 Separations 08 00052 i002199.251001.872.432.433.440.121
Tebuconazole
(TEB)
C16H22ClN3O Separations 08 00052 i003307.820003.653.693.692.010.036
a ChemAxon (http://www.chemicalize.org, 2 January 2021). b Hazardous Substances Data Bank (http://pubchem.ncbi.nih.gov, 2 January 2021).
Table 2. Physicochemical characteristics of PSB and PSBOX-A.
Table 2. Physicochemical characteristics of PSB and PSBOX-A.
PropertiesPSBPSBOX-A
Bulk elemental
constitution (%)
C86.682.1
H1.61.1
O3.74.9
N1.71.3
S0.10.2
Ash6.310.5
Atomic ratioH/C0.20.1
O/C0.030.04
N/C0.020.01
Specific surface area (m2 g−1)93.91977.6
Total pore volume (cm3 g−1)0.040.12
Average pore size (nm)3.83.4
Table 3. Kinetic parameters for the adsorptions of the selected fungicides using PSB and PSBOX-A (agitation speed = 150 rpm; contact time = 24 h; absorbent dosage = 50 mg L−1; initial concentration of each fungicide = 10 μM; temperature = 25 °C; pH = 7.0).
Table 3. Kinetic parameters for the adsorptions of the selected fungicides using PSB and PSBOX-A (agitation speed = 150 rpm; contact time = 24 h; absorbent dosage = 50 mg L−1; initial concentration of each fungicide = 10 μM; temperature = 25 °C; pH = 7.0).
AbsorbentsCompoundsQe,exp
(μmol g−1)
Pseudo-Fist-OrderPseudo-Second-Order
Qe,cal
(μmol g−1)
k1
(1 h−1)
R2Qe,cal
(μmol g−1)
k2
(g μmol−1·h)
R2
PSBCAR15.1 ± 0.16.1 ± 0.30.2 ± 0.040.97615.4 ± 0.10.09 ± 0.010.999
PYR6.1 ± 0.12.1 ± 0.30.1 ± 0.020.9156.7 ± 0.050.08 ± 0.010.999
TEB11.6 ± 0.25.4 ± 0.60.1 ± 0.030.88210.1 ± 0.10.2 ± 0.0050.999
PSBOX-ACAR196.9 ± 0.281.9 ± 1.10.3 ± 0.030.983202.0 ± 0.070.008 ± 0.0040.999
PYR179.7 ± 0.0681.4 ± 1.30.2 ± 0.020.982184.9 ± 1.00.007 ± 0.0030.999
TEB186.2 ± 0.288.3 ± 2.20.2 ± 0.050.985192.0 ± 1.60.007 ± 0.0020.999
Table 4. Isotherm parameters for the adsorptions of fungicides using PSB and PSBOX-A (agitation speed = 150 rpm; contact time = 24 h; absorbent dosage = 50 mg L−1; initial concentration of each fungicide = 10 μM; temperature = 25 °C; pH = 7.0).
Table 4. Isotherm parameters for the adsorptions of fungicides using PSB and PSBOX-A (agitation speed = 150 rpm; contact time = 24 h; absorbent dosage = 50 mg L−1; initial concentration of each fungicide = 10 μM; temperature = 25 °C; pH = 7.0).
AbsorbentsCompoundsLangmuirFreundlich
Qmax
(μmol g−1)
KL
(L μmol−1)
R2nKF
(μmol1−(1/n) L1/n g−1)
R2
PSBCAR92.6 ± 1.30.02 ± 0.0040.9991.4 ± 0.033.1 ± 0.030.995
PYR61.7 ± 2.20.03 ± 0.0010.9981.5 ± 0.042.9 ± 0.010.990
TEB66.7 ± 0.50.03 ± 0.0020.9961.5 ± 0.023.0 ± 0.050.988
PSBOX-ACAR531.2 ± 1.91.2 ± 0.10.9974.9 ± 0.07288.9 ± 0.70.764
PYR467.7 ± 1.20.8 ± 0.050.9993.9 ± 0.03214.9 ± 0.80.800
TEB495.1 ± 2.50.9 ± 0.30.9994.1 ± 0.04239.1 ± 0.60.826
Table 5. Thermodynamic parameters of fungicides adsorption onto PSB and PSBOX-A (agitation speed = 150 rpm; contact time = 24 h; absorbent dosage = 50 mg L−1; initial concentration of each fungicide = 10 μM; pH = 7.0).
Table 5. Thermodynamic parameters of fungicides adsorption onto PSB and PSBOX-A (agitation speed = 150 rpm; contact time = 24 h; absorbent dosage = 50 mg L−1; initial concentration of each fungicide = 10 μM; pH = 7.0).
AdsorbentsFungicidesTemperature
(K)
Thermodynamic Parameters
G°
(kJ mol−1)
ΔH°
(kJ mol−1)
ΔS°
(J mol−1 K)
PSBCAR288–16.6 ± 0.40.003 ± 0.040.048 ± 0.003
298–18.3 ± 0.1
308–19.5 ± 0.1
PYR288–12.8 ± 0.10.002 ± 0.00030.037 ± 0.001
298–15.2 ± 0.2
308–16.7 ± 0.1
TEB288–15.2 ± 0.10.002 ± 0.00040.045 ± 0.005
298–17.0 ± 0.2
308–18.1 ± 0.2
PSBOX-ACAR288–31.9 ± 0.10.003 ± 0.00050.063 ± 0.002
298–33.5 ± 0.2
308–35.8 ± 0.3
PYR288–28.7 ± 0.30.005 ± 0.00070.090 ± 0.011
298–30.1 ± 0.9
308–31.6 ± 0.2
TEB288–30.0 ± 0.90.005 ± 0.00060.096 ± 0.013
298–31.5 ± 0.2
308–33.0 ± 0.1
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lee, Y.-G.; Shin, J.; Kwak, J.; Kim, S.; Son, C.; Kim, G.-Y.; Lee, C.-H.; Chon, K. Enhanced Adsorption Capacities of Fungicides Using Peanut Shell Biochar via Successive Chemical Modification with KMnO4 and KOH. Separations 2021, 8, 52. https://0-doi-org.brum.beds.ac.uk/10.3390/separations8040052

AMA Style

Lee Y-G, Shin J, Kwak J, Kim S, Son C, Kim G-Y, Lee C-H, Chon K. Enhanced Adsorption Capacities of Fungicides Using Peanut Shell Biochar via Successive Chemical Modification with KMnO4 and KOH. Separations. 2021; 8(4):52. https://0-doi-org.brum.beds.ac.uk/10.3390/separations8040052

Chicago/Turabian Style

Lee, Yong-Gu, Jaegwan Shin, Jinwoo Kwak, Sangwon Kim, Changgil Son, Geon-Youb Kim, Chang-Ha Lee, and Kangmin Chon. 2021. "Enhanced Adsorption Capacities of Fungicides Using Peanut Shell Biochar via Successive Chemical Modification with KMnO4 and KOH" Separations 8, no. 4: 52. https://0-doi-org.brum.beds.ac.uk/10.3390/separations8040052

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