Next Article in Journal
Changes in Extreme Precipitation on the Tibetan Plateau and Its Surroundings: Trends, Patterns, and Relationship with Ocean Oscillation Factors
Previous Article in Journal
The Sedimentation Rate in the Crimean Hypersaline Lake Aktashskoye Estimated Using the Post-Chernobyl Artificial Radionuclide 90Sr as a Radiotracer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Preparation of MnOx-Modified Biochar and Its Removal Mechanism for Cr(VI) in Aqueous Solution

School of River and Ocean Engineering, Chongqing Jiaotong University, Chongqing 400074, China
*
Author to whom correspondence should be addressed.
Water 2022, 14(16), 2507; https://doi.org/10.3390/w14162507
Submission received: 24 July 2022 / Revised: 4 August 2022 / Accepted: 11 August 2022 / Published: 14 August 2022
(This article belongs to the Section Wastewater Treatment and Reuse)

Abstract

:
Biochar (BC) is considered to be a great potential adsorbent to remove various contaminants, but the sorption capacity for chromium (Cr) is predominantly limited for the net negative charge. In this study, BC from rice husk was impregnated with Mn(NO3)2 solution to synthesize MnOx-modified BC (MnOBCs) for enhancing Cr(VI) removal in an aqueous solution. MnOBCs were characterized, and MnOx (manganese ore) was found to be the dominant crystal in MnOBCs. Batch sorption and kinetic experiments combined with spectral analysis were carried out to elucidate the sorption capacity and mechanisms of Cr(VI) sorption onto BC and MnOBCs. Results showed that the sorption kinetic process fitted to the Elovich model, and the modification enhanced the sorption capacity of Cr(VI) on BC. Compared to ion strength, pH is the main control factor for Cr(VI) fixed on BC and MnOBCs, and the sorption amount decreased with the pH value increasing. Moreover, X-ray photoelectron spectroscopy results showed that the proportion of Mn(II) decreased from 33.59% to 8.33%, and that of Mn(VI) increased from 30.58% to 52.72% after Cr(VI) sorption on MnOBCs. Meanwhile, the reduction reaction of Cr(VI) occurred during the sorption process on the BC and MnOBCs surface, and the reduction reaction was more obvious on the MnOBCs surface. MnOx loaded in BC can enhance the electrostatic attraction and redox capacity, which can improve the removal of Cr(VI) in an aqueous solution. This study provides information on the sorption and redox of Cr(VI) on BC, and allows us to better understand the mechanism of Cr(VI) removal in solutions by MnOBCs.

Graphical Abstract

1. Introduction

Chromium (Cr) is a strong carcinogen and frequently detected in water and soil environments owing to natural and anthropogenic activities, including tanning, dyeing, chemical manufacturing, and metal smelting [1]. Cr(VI) and Cr(III) are dominant forms of Cr in water, and they can interconvert with each other in the environment [2]. Cr(III) is an important component of the glucose tolerance factor and participates in lipid metabolism regulation [3]. However, Cr(VI) compounds have attracted attention because of their potential toxic effects on humans, animals, plants, and microorganisms [4]. Thus, it is essential to further explore the removal method of Cr(VI) in the environment by its transformation and immobilization.
Currently, many Cr(VI) removal methods have been studied, such as biological reduction [5], chemical precipitation [6], sorption [7,8], complexation [9], and photocatalytic removal [10]. In addition, sorption is considered a practical method for Cr removal from wastewater due to its wide applicability and high efficiency [11]. Many carbon-based materials and their derivatives (such as activated carbon [12], carbon nanotubes [13], graphene [14], and BC [15]) have been used used to remove Cr(VI).
Biochar (BC) is a product of high-temperature pyrolysis of carbon-rich biomass, which is obtained from the waste of agriculture [16], forestry [17], animal husbandry [18], and fishery [19], under anaerobic conditions. Meanwhile, BC has been considered as a new way to increase the “carbon sink” [20]. In recent years, BC was used as an excellent adsorbent to remove hazardous chemical compounds because of its special physicochemical properties [21,22,23]. For example, BC showed good sorption for Pb and Cd, and maximum sorption capacities reached 359 mg g−1 and 135.7 mg g−1, respectively [24,25]. Biochar from freshwater hornwort and macroalgae showed the same good sorption capacity for Cr(III), with maximum sorption of 163.9 mg g−1 and 60.2 mg g−1, respectively [26]. It was also shown that the sorption of Cr(III) by freshwater macroalga biochar increased with the increase of pyrolysis temperature [27]. Meanwhile, the maximum adsorption capacity of Cr(III) by biochar pyrolyzed from waste tomato leaves and stems at 800 °C was 169.5 mg g−1 [28]. However, BC is not good at removing anionic negative ions, such as Cr(VI), As, Se, and Sb. For instance, pyrolytic BC from bean straw was difficult to adsorb Cr(VI), and sorption capacity was only 6.48 mg g−1 [29]. The maximum sorption capacity of As on BC prepared from pine wood was only 0.265 mg g−1 [30]. BC pyrolyzed from pine birch and bamboo at 600 °C for 1 h showed a sorption capacity of 2.59 and 3.35 mg g−1 for Cr(VI), respectively [31].
Nevertheless, some studies expressed that BC could be modified by other materials to improve its removal capacity for contaminants in an aqueous solution. The physical, chemical, and biological methods used in BC modification can enhance its sorption efficiency [32]. Unlike other physical and chemical modification methods, metal oxide loading onto BC can increase the positive charge and its affinity for negatively charged metals [33]. Modified magnetic biochar (BC) was prepared from the litchi shell for removing Cr(VI) from wastewater, and this adsorbent is not only reusable but also easy to recycle [34]. For example, Fe-Mn-Ce oxide-modified corn stover BC composites enhanced the electrostatic attraction between the composites and As(III) [35]. Similarly, the Cr(VI) anion interacts electrostatically on the surface of the positively charged Fe-Mn oxide-modified BC [36]. On the other hand, copper-modified activated bamboo charcoal (MAC) can have more effective removal rates for Cd, Pb, As, and Cr over a wider pH range than commercial activated charcoal (CAC) [37]. Meanwhile, Cu-modified activated bamboo charcoal (10BCKOH) was used to remove Ru dyes from an aqueous solution mainly by the ion-exchange mechanism [38]. As shown in Table 1, the maximum adsorption capacity of MnOBCs for Cr(VI) was higher than that of most other modified biochars in the literature reported in recent years. Meanwhile, most studies have focused on Fe-modified biochar. Fewer studies have been conducted on Cr(VI) removal by Mn-modified biochar.
Manganese oxides are important minerals in soils with many advantages, including variability, abundant hydroxyl groups, and often applied to remove contamination [39]. Nevertheless, the high cost of MnOx limits its utilization as a sorbent. The combination of BC and nanomaterials promotes its efficiency in environmental remediation [40]. Therefore, MnOx-modified BC was used for pollutant removal with favorable performance [41]. For example, MnOx-modified BC was prepared by subjecting BC to a series of reactions with KMnO4 and exhibited excellent sorption capacity for CuCA [42]. In addition, manganese oxide-modified macro-mesoporous BC was prepared and showed a strong preference for heavy metal sorption when high concentrations of the substances coexisted [43]. On the other hand, using the high reduction capacity of MnOx, the manganese iron oxide-modified corn stover BC reduced Cr(VI) to Cr(III) effectively, and reduced the toxicity of Cr [36].
Considering these findings, MnOx plays a very important role in BC modification and Cr(VI) reduction, but it remains unclear what the valence state of Mn exerts on such an electron transfer process with BC. Therefore, in this study, (1) BC was prepared from rice husk and modified by MnOx using chemical impregnation; (2) The physicochemical property of BC and modification of BC were determined and characterized; (3) Sorption kinetic and isothermal sorption experiments of Cr(VI) were carried out to obtain the sorption capacity, and the effects of pH and ion strength on the removal of Cr(VI) were investigated. (4) Energy-dispersive spectroscopy (EDS) and X-ray photoelectron spectroscopy (XPS) were analyzed to elucidate the removal mechanism of Cr(VI) by BC and modified BC.

2. Materials and Methods

2.1. Chemical and Reagents

All chemicals were analytical grade or better. Potassium dichromate (K2Cr2O7), phosphoric acid (H3PO4), and manganous nitrate (Mn(NO3)2) were from Chengdu Kolon Chemical Co., Chengdu, China. Sodium hydroxide (NaOH), sulfuric acid (H2SO4), hydrochloric acid (HCl), Nitric acid(HNO3), and acetone(C3H6O) were from Chongqing Chuandong Chemical Group Co. Chengdu, China. Diphenylcarbazide(C13H14N4O) was from Shanghai Aladdin Biochemical Technology Co., Ltd. Shanghai, China. The solvent used in experiments was deionized water.

2.2. The Modification of BC

Rice husk (a typical biomass waste) collected from a farmland in Chongqing, China, was used to prepare BC at 600 °C. Briefly, rice husk was dried and ground to less than 2.0 mm and then carbonized in a tube furnace at 600 °C with 5.0 °C min−1 under N2 protection for 2.0 h. N2 flow rate of 200 mL min1. The prepared BC was ground to pass through a 0.15 mm sieve size, and then placed in a ziplock bag and set aside.
Then, BC was immersed in HNO3 solution with a mass fraction of 10% for 1.5 h, and washed with deionized water to neutral and finally dried in an air oven. The preparation process of MnOx-modified BC (MnOBCs) is shown in Figure 1. First, BC (after acid-washed) was impregnated with 0.05 or 0.1 mol L−1 Mn(NO3)2 solution for 6 h according to a solid–liquid ratio of 1:10. Next, the water was evaporated in a constant temperature water bath and then dried in a vacuum drying oven at 90 °C for 24 h. Finally, the dried BC was placed into a tube furnace under N2 protection and baked at 600 °C for 3 h. It was processed twice following the process described above. MnOBCs was then obtained. The samples impregnated with 0.05 mol L−1 Mn(NO3)2 were named MnOBC-1. The samples impregnated with 0.1 mol L−1 Mn(NO3)2 were named MnOBC-2, whereas the untreated BC was labeled as BC.

2.3. Physicochemical Characterization

Fourier transform infrared spectrometer (Nicolet-iS10, Thermo Fisher Scientific, Waltham, MA, USA) was used to obtain the surface functional groups of BC and MnOBCs. The surface morphology was examined by scanning electron microscope (SEM, Quanta 450 FEG, FEI, Hillsboro, OR, USA) and transmission electron microscope (TEM, Talos F200X, Thermo Fisher Scientific, MA, USA). The elemental composition in the specific area was examined by Energy Dispersive Spectrometer (EDS, SuperX, FEI, OR, USA). The specific surface area and pore space characterization were analyzed by the multi-channel specific surface area and pore size analyzer (MicroActive for TriStar II Plus 2.02, Micromeritics Instrument Corporation, Norcross, GA, USA). The crystal structure was analyzed by X-ray diffractometry (Bruker D8, Bruker Corporation, Billerica, MA, USA). The chemical state and surface composition of the MnOBCs were determined by X-ray photoelectron spectrometer (XPS, Escalab 250Xi, Thermo Fisher Scientific, MA, USA). Cr(VI) concentration was measured by spectrophotometry with diphenylcarbazide. In the diphenylcarbazide spectrophotometric method, the color developer was prepared by dissolving 0.2 g of diphenylcarbazide in 50 mL of acetone solvent and diluting it to 100 mL with water. To determine the concentration of Cr(VI) in the solution, 0.5 mL (1 + 1) sulfuric acid solution, 0.5 mL (1 + 1) phosphoric acid, and 2 mL color developer were added sequentially to 50 mL of the sample solution. After the color development reaction at 5–10 min, the resulting mixture was measured with a UV-Vis spectrophotometer (UV-1800, Aoyi Instruments (Shanghai) Co. Shanghai, China) at the wavelength of 540 nm.

2.4. Sorption Experiments

K2Cr2O7 was dissolved in deionized water to gain the Cr(VI) stock solution, and the pH was adjusted by 0.1 mol L−1 HCl and 0.1 mol L−1 NaOH. The influence of the initial pH on Cr(VI) removal was investigated by mixing 0.025 g of BC or MnOBCs and 25 mL Cr(VI) solution (50 mg L−1), and the initial pH ranged from 2.0 to 10.0. Then, suspensions were shaken 24 h at 150 rpm.
The influence of ionic strength on the removal of Cr(VI) by BC and MnOBCs was investigated. BC or MnOBCs (1 g L−1) was mixed with 50 mg L−1 Cr(VI) solution, and the concentration of NaCl was designed to 0, 0.1, 1, and 10 mmol L−1, and the pH was adjusted to 2. Then, it was shaken at 25 °C and 150 rpm for 24 h.
Kinetic sorption experiments were performed in batch tests with the 50 mg L−1 Cr(VI) solution. First, 0.1 g of BC or MnOBCs was weighed in a conical flask and added to the 100 mL of Cr(VI) solution. Then, the pH value was adjusted to 2, and the samples were continuously shaken in a water bath with an oscillation frequency of 150 rpm for 48 h. At the designed time, a certain amount of suspension was taken out. The Cr(VI) concentration in the filter liquid was measured. Batch sorption experiments were carried out, and initial concentration of Cr(VI) was from 0 mg L−1 to 200 mg L−1, and the pH value was adjusted to 2. The suspension was shaken for 48 h.
All samples were filtered through a 0.45 μm filter membrane, and the Cr(VI) concentration in the filter liquor was tested. In addition, all experiments were repeated twice, and the average value of the results from the analysis was taken.

2.5. Data Analysis

The sorption capacity of Cr(VI) on BC or MnOBCs was calculated by the following Equation (1),
q e = V   ( C 0     C e ) M
where qe (mg g−1) is the sorption capacity, V (L) is the volume, C0 and Ce (mg L−1) are the initial and equilibrium Cr(VI) concentrations, respectively, and M (g) is the adsorbent weight.
To evaluate the performance of different adsorbent materials, three different models, including the pseudo-first-order model (Equation (2)), pseudo-second-order model (Equation (3)), and Elovich model (Equation (4)) [54], were applied to describe the sorption kinetics of different sorbent systems.
In(qe − qt) = Inqe − k1t
t q t = 1 k 2 q e 2 + t q e
q t = 1 β   In ( α β ) + 1 β   In ( t )
where qt and qe (mg g−1) were the sorption capacity at time t and equilibrium, respectively. k1 (1 min−1) and k2 (g mg−1 min−1) were the sorption rate constant, α (mg g−1 min−1) and β (g mg−1) are Elovich constants, which denotes the initial sorption rate constant and desorption rate constant, respectively.
The Langmuir and Freundlich isotherm models are widely used to study the sorption mechanism, and the equations are given as follow.
q e = q max K L C e 1   +   K L C e
q e = K F C e 1 / n
where Ce (mg L−1) is the equilibrium concentration; qmax (mg g−1) is the maximum sorption capacity; KL (L mg−1), KF (mg1−1/n L1/n g−1), and n (dimensionless) are the isotherm constants.
Experimental data were processed with Excel software, and curves were drawn with Origin 2021 software. Solution spectrum analysis and material crystallography were performed using XPSPEAK and Jade.

3. Results and Discussion

3.1. The Physiochemical Properties of BC and MnOBCs

3.1.1. Specific Surface Area and Pore Structure

N2 adsorption–desorption isotherms of BC and MnOBCs samples are given in Figure S1. The isotherms of all specimens belonged to type IV, and the isotherms had hysteresis loops (H3 for BC and H4 for MnOBCs), indicating that they presented mesoporous structures [55]. The pore size distribution (Figure S1) of MnOBCs ranged from 2 nm to 10 nm, suggesting that they are mesoporous materials [56]. Meanwhile, the specific surface areas (Table 2) of MnOBCs were nearly 10 times larger than those of BC, which may be due to the dispersed attachment of MnOx particles loaded on MnOBCs. In addition, the average pore size of MnOBCs decreased with the enrichment of MnOx, which may be due to the formation of MnOx and oxygen-containing functional groups blocking the mesopores and micropores of MnOBCs.

3.1.2. Surface Morphology Characteristics

Figure 2 presents the morphological characteristics of BC and MnOBCs samples. We can see that the BC had clear pores, a honeycomb network, and different shapes of ash particles that are attached. The surface morphology of MnOBCs was rough, and nanoparticles were adsorbed on their surface and pores. As the concentration of Mn(NO3)2 increased, the coverage area of the particles attached to the surface increased (Figure 2b,c). These results confirmed that MnOx loaded on the BC successfully. Of note, the substrate surface of MnOBC-2 was mostly covered by MnOx.

3.1.3. Analysis of Functional Groups and Mineralogical Characterization

The FT-IR spectra (Figure 3a) showed that BC and MnOBCs were rich in functional groups. For instance, there was an intense peak at 3429 cm−1, due to the stretching vibration of the O-H bond in the hydroxyl group [57]. There was a spectral band at about 1574 cm−1, due to C=O and C=C bonds [58]. Moreover, the high-intensity peak around 1094 cm−1 corresponded to C-O stretching and C-OH bending, whereas the peak at 797 cm−1 represented the asymmetric bending vibration of Si-O-Si [59,60]. The peak at 1417 cm−1 represented the stretching vibration of the carboxyl group (-COOH). However, the characteristic peaks of MnOBCs at 3429 cm−1 (hydroxyl vibration peak) became weaker; thus, the absorption peak at 797 cm−1 also became weaker, indicating that the ash on the BC was cleaned during the pre-preparation process. Abundant functional groups in BC can provide reaction or interaction sites for Cr(VI) transformation and sorption on the material surface [61,62].
XRD was performed to characterize the crystallinity of the composites. As seen from the XRD spectrum (Figure 3b), BC had poor crystallinity, and both BC and MnOBCs showed broad diffraction peaks at 2θ = (20–30)°, which were diffraction peaks of an amorphous structure. With the help of Jade 6 software, the characteristic peaks of MnOBCs were located at 2θ = 35.1°, 40.7°, 58.9°, 70.4°, 74.0°, and 88.1°, which are in agreement with the standard card spectra (PDF#-07-0230), corresponding to the 111, 200, 220, 311, 222, and 400 crystal planes of crystalline manganese ore, respectively [63]. Combined with the SEM images of MnOBCs (Figure 2), it was confirmed that MnOx was successfully loaded onto the BC, and the impregnation of MnOx on the surface of BC was directly related to the Mn(NO3)2 concentration.

3.2. Cr(VI) Sorption Performance of BC and MnOBCs

3.2.1. Effect of Initial pH on Cr(VI) Sorption

In general, the Cr(VI) form and surface charge properties of the material should be influenced by pH. In this study, five pH gradients from acidic to basic (pH = 2, 4, 6, 8, 10) were selected to investigate the effect of pH on the sorption amount of Cr(VI) on BC and MnOBCs. Results indicated that sorption amount of Cr(VI) decreased with increasing pH (Figure 4a). This phenomenon is attributed to the formation of Cr(VI) changed with pH, and Cr(VI) in the solution exists in the form of HCrO4 at pH ≤ 3. HCrO4 can be converted to Cr2O72− and CrO42− when the pH increases, and CrO42− plays a dominant role if the pH is greater than 6.0 [2]. In addition, the sorption free energy of HCrO4 is lower than that of Cr2O72− and CrO42− [64], which may lead to the higher sorption amount of Cr(VI) in a more acidic condition. Meanwhile, MnOBCs usually exhibit a high zero charge point (pHpzc) [65]. At lower pH values, MnOBCs have positively charged surfaces and interact more readily with negatively charged ions. Meanwhile, the protonation on the BC and MnOBCs surface gradually decreases with pH increasing, and the sorption capacity is weakened [66]. In addition, the reduction reaction of Cr(VI) to Cr(III) is a proton-consuming process, which could be inhibited with pH increasing [67]. In one word, both of the surface charge of MnOBCs and the intensity of Cr(VI) reduction reaction varied with pH, which can lead to the different sorption capacity at different pH [68,69].

3.2.2. Effect of Ion Strength on Cr(VI) Sorption

The ions concentration in the sorption system may affect the properties of sorption [70]. The ions produced during the hydrolysis of these salts may reduce the sorption efficiency of metal ions through competing sorption sites [71]. To better simulate the actual water conditions, the effect of NaCl concentrations on the removal of Cr(VI) by BC and MnOBCs was investigated at pH = 2.0. As shown in Figure 4b, the sorption amount of Cr(VI) on BC and MnOBCs did not change clearly when the ionic strength increased. It was suggested that Cr(VI) removal by BC and MnOBCs cannot be affected by the ion strength.

3.2.3. Sorption Kinetics

To further investigate the sorption mechanism, sorption kinetic experiments of Cr(VI) sorption on BC and MnOBCs were conducted. As shown in Figure 5, the amount of Cr(VI) sorption on BC and MnOBCs increased with prolonged reaction time. Sorption processes seemed quick at the beginning, and slowed down after 6 h, because of most sorption sites occupied and the resistance enhanced [72]. The pseudo-first-order and pseudo-second-order model were employed to fit the sorption kinetic data, and the correlation coefficients (Table 3) changed in the range of 0.7281–0.8570 and 0.8429–0.9334, respectively. It was indicated that the experimental data did not conform to the pseudo-first-order and pseudo-second-order kinetic models. However, the Elovich equation fitted for sorption process properly, and the correlation coefficients were 0.9799 to 0.9821. The Elovich equation is often used to describe the chemisorption kinetics on highly heterogeneous adsorbents [73]. It indicates that the sorption process of MnOBCs with respect to Cr(VI) was mainly via chemisorption. The higher α values than β values may also reflect the rapid sorption in the early stages [74].

3.2.4. Sorption Isotherms

The sorption isotherms of Cr(VI) on BC and MnOBCs are shown in Figure 6. The Langmuir and Freundlich models were used to understand the sorption mechanism. The Langmuir model was derived from the hypothesis of monolayer sorption, and the Freundlich model was used for multilayer sorption [75]. The correlation coefficient values (Table 4) obtained from the Langmuir and Freundlich equilibrium isotherm models show that the Freundlich model could better interpret the sorption process of Cr(VI) by MnOBCs, indicating that the sorption of Cr(VI) on MnOBCs was a heterogeneous chemisorption process. The sorption capacity of BC for Cr(VI) was very weak, and the maximum sorption capacity fitted by the Langmuir model was only 10.10 mg g−1. The maximum sorption capacities of MnOBC-1 and MnOBC-2 for Cr(VI) were 20.69 and 28.58 mg g−1, which are 2.04 and 2.83 times higher than those of BC, respectively. Combined with the SEM plots (Figure 2), it further illustrates that the higher the loaded MnOx content, the higher the sorption amount.
For the Freundlich equation, the value of 1/n usually ranges from 0 to 1, representing the exchange strength or surface inhomogeneity [76]. The larger the value of the sorption equilibrium constant KF, the greater the sorption capacity [77]. As shown in Table 4, MnOBC-2 had the largest KF value, indicating that it had the largest sorption capacity. The sorption capacity of BC was minimal. The maximum sorption capacity of MnOBC-2 for Cr(VI) might be related to electrostatic interaction and the higher specific surface area for the largest amount of MnOx loaded. In addition, the Mn-O-Cr complexation might lead to more sorption capacity in MnOBCs.

3.3. Formation and Distribution of Cr in BC and MnOBCs

In general, Cr(III) is less toxic than Cr(VI) and less mobile due to the limited solubility of the hydroxide [66]. Therefore, the immobilization mechanism of Cr(VI) on BC and MnOBCs was determined by XPS, EDS.

3.3.1. XPS Spectra Analysis

In this work, Cr species in BC and MnOBCs was determined by XPS, which can be applied to analyze the surface chemical composition and oxidation states. Typically, the binding energy of the Cr2p3/2 at 576.9 eV represents Cr(III), and the binding energy at 578.0 eV represents Cr(VI) [78]. As shown in Figure 7, Cr(VI) and Cr(III) were found on the surface of BC and MnOBCs. This may be because Cr(VI) was immobilized on the surface of the material due to the texture and highly porous surface of BC, and then the free radical groups on the surface of BC reduced some of the adsorbed Cr(VI) to Cr(III) [79]. In addition, the proportion of Cr(III) on the surfaces of MnOBC-1 and MnOBC-2 was 32.45% and 44.61%, respectively, which were larger than that of Cr(III) on the surface of BC (28.45%). The difference in the reduction capacity of Cr(VI) indicated the difference between the reduction components of BC and MnOBCs, further suggesting that Mn elements in MnOBCs were also involved in the Cr(VI) reduction reaction.
To provide insights into the Cr(VI) reduction and sorption mechanisms, XPS analysis was performed on the MnOBC-2 with and without Cr(VI) sorption, and shown in Figure 8. As Figure 8a shown, elemental Mn peaks were found in XPS spectroscopy of MnOBCs, and the Mn2p3/2 spectrum (Figure 8b) had three peaks at binding energies of 641.4, 642.3, and 643.5 eV, which were attributed to Mn(II), Mn(III), and Mn(IV), respectively [80]. The valence of Mn elements on the surface of MnOBC-2 was mainly composed of Mn(II), Mn(III), and Mn(IV), accounting for 33.59%, 35.83%, and 30.58%, respectively. However, in the MnOBC-2 with Cr(VI) sorption, the percentage of Mn(II) decreased to 8.33%, whereas the percentage of Mn(III) and Mn(IV) increased to 38.95% and 52.72%, respectively. This indicates that during the sorption of Cr(VI), the Mn on the surface of MnOBC-2 was oxidized, as Mn(II) was mainly oxidized to Mn(III) and Mn(IV). Therefore, the removal mechanism of Cr(VI) also includes redox reactions. The negatively charged Cr(VI) migrates to the positively charged MnOBCs surface under electrostatic interaction. The fixed portion of Cr(VI) was reduced to Cr(III) by BC-loaded MnOx and free radicals and their reduction, whereas Mn(II) in MnOx was oxidized and its valence state was elevated.
To further elucidate the mechanism of Cr(VI) sorption, Figure 8c shows the C-functional group distribution on MnOBC-2. As shown in Figure 8c, the C1s spectra of MnOBC-2 represented C-C, C-O, and O=C-O characteristic peaks at 284.8, 286.1, and 288.7 eV, respectively [81]. The area percentages of C-O and O=C-O of MnOBC-2 were 38.01% and 12.45%, respectively. In contract, after the sorption of Cr(VI), these percentages decreased to 19.40% and 10.64%, respectively. It is indicated that the O=C-OH functional group plays a key role in the removal of Cr(VI), which can be attributed to the complexation reaction between the carboxyl group (COOH) and Cr(VI). The O1s spectra (Figure 8d) showed that the surface of MnOBC-2 contained not only O from the BC itself but also O from MnOx. The binding energies 530.1, 531.4, 532.2, and 533.4 eV were chosen as the splitting peak positions corresponding to Mn-O (the lattice oxygen), Mn-OH, C-OH, and O-C=O [82,83]. Compared to MnOBC-2, the percentage of the peak area of Mn-OH decreased from 32.05% to 12.71% when Cr(VI) was sorption on MnOBC-2, indicating that the hydroxyl group was involved in the Cr(VI) sorption. On the contrary, the area of the peak assigned to Mn-O increased from 2.59% to 9.23%, probably forming Mn-O-Cr on the surface of MnOBC-2.

3.3.2. TEM-EDS Analysis

TEM-EDS analysis was performed for BC and MnOBCs with Cr(VI) sorption to understand the composition and distribution of elements (i.e., C, O, Ca, Si, Mn, and Cr) (Figure 9). Therefore, the immobilization mechanism of Cr(VI) on BC and MnOBCs was deter-mined by XPS and TEM-EDS. As shown in Figure 9, the EDS indicated that Cr was detected in BC and MnOBCs. In addition, the presence of Mn elements was detected in MnOBC-2, which further indicates that MnOx was successfully loaded on BC. Meanwhile, the surface of MnOBCs contains not only C, O, and Mn but also K, Ca, Na, Mg, S, and P, which were present in BC as essential elements for growth and affected the fixation of Cr(VI) in the aqueous solution by BC.

3.4. Cr(VI) Immobilization Mechanisms on BC and MnOBCs

The impregnation of BC with MnOx could improve its surface morphology, functional group modification, and elemental composition, and results in the sorption capacity of the Cr(VI) increase [84]. In addition, the MnOBCs exhibited a higher reduction potential and ability to fix Cr(VI) [85]. There are several possible pathways for the immobilization mechanism of Cr(VI) on BC and MnOBCs: (1) Compared with BC, MnOBCs have a larger specific surface area and more sorption sites, which could fix more Cr(VI). (2) Hydroxyl (-OH) and carboxyl (-COOH) groups could be involved in the immobilization of Cr(VI), and Mn-O-Cr was formed on the surface of MnOBCs, which is given in the XPS analysis results. The surface of BC was rich in carboxyl and hydroxyl groups, which had strong affinity for Cr(VI) ions [86]. (3) With the loading of MnOx, the positive charge on the BC surface increases, facilitating the fixation of Cr(VI) by electrostatic attraction. Previous studies demonstrated that manganese oxides had a positively charged mineral surface [87]. (4) The loading of MnOx increases the reduction of Cr, whereas Cr(III) is more easily immobilized. Meanwhile, previous studies demonstrated that when Mn(II) and Cr(VI) are adsorbed in ferrihydrite at the same time, Mn(II) and Cr(VI) undergo redox reactions [88]. Therefore, BC-loaded MnOx particles may play an important role in the immobilization of Cr(VI) in the solution.

4. Conclusions

In this work, MnOBCs were successfully prepared by the chemical impregnation method to improve the sorption of Cr(VI). The specific surface area of MnOBCs increased by nearly 10 times than that of BC. The sorption kinetic process of Cr(VI) could be fitted by the Elovich kinetic model, and sorption isotherm data could be fitted by the Freundlich equation. In addition, the sorption capacity of Cr(VI) decreased sharply with the increase in pH and could not be influenced by ionic strength. Moreover, Cr(VI) and Cr(III) were found on the surface of BC and MnOBCs, and Mn(II) on the surface of MnOBCs was oxidized to Mn(III) and Mn(IV) during the Cr(VI) sorption process. Meanwhile, the functional groups (-COOH and -OH) on BC substrates were also involved in the Cr(VI) sorption by complexation reaction. In short, the sorption mechanism of Cr(VI) on MnOBCs includes the electrostatic attraction, complexation, and reduction reaction of Cr(VI). This study sheds light on the change of Mn valence in MnOBCs during Cr(VI) immobilization, and MnOBCs are potentially promising for environmental applications in the removal of Cr(VI) from an aqueous solution due to their low cost and effective sorption capacity.

Supplementary Materials

The following supporting information can be downloaded at: www.mdpi.com/article/10.3390/w14162507/s1, Figure S1: N2 adsorption–desorption isotherms and the corresponding BJH desorption pore width distribution for the BC (a), MnOBC-1 (b), and MnOBC-2 (c).

Author Contributions

All authors contributed to the study conception and design. Material preparation, data collection and analysis were performed by J.F., L.Q., T.D., Z.Q., and L.Z. The first draft of the manuscript was written by J.F. and all authors commented on previous versions of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The authors declare that all relevant data supporting the findings of this study are included in this article.

Acknowledgments

This work was financially supported by the National Natural Science Foundation of China (41977337), Chongqing Liuchuang Plan Innovation Project (No. cx2021129), the Natural Science Foundation of Chongqing of China (cstc2018jcyjAX0054).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Huang, X.; Hou, X.; Song, F.; Zhao, J.; Zhang, L. Facet-Dependent Cr(VI) Adsorption of Hematite Nanocrystals. Environ. Sci. Technol. 2016, 50, 1964–1972. [Google Scholar] [CrossRef] [PubMed]
  2. Yang, Y.; Zhang, Y.; Wang, G.; Yang, Z.; Xian, J.; Yang, Y.; Li, T.; Pu, Y.; Jia, Y.; Li, Y.; et al. Adsorption and Reduction of Cr(VI) by a Novel Nanoscale FeS/Chitosan/Biochar Composite from Aqueous Solution. J. Environ. Chem. Eng. 2021, 9, 105407. [Google Scholar] [CrossRef]
  3. Lewicki, S.; Zdanowski, R.; Krzyżowska, M.; Lewicka, A.; Dębski, B.; Niemcewicz, M.; Goniewicz, M. The Role of Chromium III in the Organism and Its Possible Use in Diabetes and Obesity Treatment. Ann. Agric. Environ. Med. 2014, 21, 331–335. [Google Scholar] [CrossRef] [PubMed]
  4. Saha, R.; Nandi, R.; Saha, B. Sources and Toxicity of Hexavalent Chromium. J. Coord. Chem. 2011, 64, 1782–1806. [Google Scholar] [CrossRef]
  5. Bai, Y.-N.; Lu, Y.-Z.; Shen, N.; Lau, T.-C.; Zeng, R.J. Investigation of Cr(VI) Reduction Potential and Mechanism by Caldicellulosiruptor saccharolyticus under Glucose Fermentation Condition. J. Hazard. Mater. 2018, 344, 585–592. [Google Scholar] [CrossRef]
  6. Zhu, Y.; Li, H.; Zhang, G.; Meng, F.; Li, L.; Wu, S. Removal of Hexavalent Chromium from Aqueous Solution by Different Surface-Modified Biochars: Acid Washing, Nanoscale Zero-Valent Iron and Ferric Iron Loading. Bioresour. Technol. 2018, 261, 142–150. [Google Scholar] [CrossRef] [PubMed]
  7. Gan, C.; Liu, Y.; Tan, X.; Wang, S.; Zeng, G.; Zheng, B.; Li, T.; Jiang, Z.; Liu, W. Effect of Porous Zinc–Biochar Nanocomposites on Cr(VI) Adsorption from Aqueous Solution. RSC Adv. 2015, 5, 35107–35115. [Google Scholar] [CrossRef]
  8. Zheng, C.; Yang, Z.; Si, M.; Zhu, F.; Yang, W.; Zhao, F.; Shi, Y. Application of Biochars in the Remediation of Chromium Contamination: Fabrication, Mechanisms, and Interfering Species. J. Hazard. Mater. 2021, 407, 124376. [Google Scholar] [CrossRef]
  9. Wang, F.; Liu, L.-Y.; Liu, F.; Wang, L.-G.; Ouyang, T.; Chang, C.-T. Facile One-Step Synthesis of Magnetically Modified Biochar with Enhanced Removal Capacity for Hexavalent Chromium from Aqueous Solution. J. Taiwan Inst. Chem. Eng. 2017, 81, 414–418. [Google Scholar] [CrossRef]
  10. Chen, G.; Feng, J.; Wang, W.; Yin, Y.; Liu, H. Photocatalytic Removal of Hexavalent Chromium by Newly Designed and Highly Reductive TiO2 Nanocrystals. Water Res. 2017, 108, 383–390. [Google Scholar] [CrossRef]
  11. GracePavithra, K.; Jaikumar, V.; Kumar, P.S.; SundarRajan, P. A Review on Cleaner Strategies for Chromium Industrial Wastewater: Present Research and Future Perspective. J. Clean. Prod. 2019, 228, 580–593. [Google Scholar] [CrossRef]
  12. Chen, Y.; Qian, Y.; Ma, J.; Mao, M.; Qian, L.; An, D. New Insights into the Cooperative Adsorption Behavior of Cr(VI) and Humic Acid in Water by Powdered Activated Carbon. Sci. Total Environ. 2022, 817, 153081. [Google Scholar] [CrossRef] [PubMed]
  13. Ihsanullah; Al-Khaldi, F.A.; Abu-Sharkh, B.; Abulkibash, A.M.; Qureshi, M.I.; Laoui, T.; Atieh, M.A. Effect of Acid Modification on Adsorption of Hexavalent Chromium (Cr(VI)) from Aqueous Solution by Activated Carbon and Carbon Nanotubes. Desalin. Water Treat. 2016, 57, 7232–7244. [Google Scholar] [CrossRef]
  14. Dubey, R.; Bajpai, J.; Bajpai, A.K. Green Synthesis of Graphene Sand Composite (GSC) as Novel Adsorbent for Efficient Removal of Cr (VI) Ions from Aqueous Solution. J. Water Process Eng. 2015, 5, 83–94. [Google Scholar] [CrossRef]
  15. Zhong, M.; Li, M.; Tan, B.; Gao, B.; Qiu, Y.; Wei, X.; Hao, H.; Xia, Z.; Zhang, Q. Investigations of Cr(VI) Removal by Millet Bran Biochar Modified with Inorganic Compounds: Momentous Role of Additional Lactate. Sci. Total Environ. 2021, 793, 148098. [Google Scholar] [CrossRef]
  16. Qu, J.; Wang, Y.; Tian, X.; Jiang, Z.; Deng, F.; Tao, Y.; Jiang, Q.; Wang, L.; Zhang, Y. KOH-Activated Porous Biochar with High Specific Surface Area for Adsorptive Removal of Chromium (VI) and Naphthalene from Water: Affecting Factors, Mechanisms and Reusability Exploration. J. Hazard. Mater. 2021, 401, 123292. [Google Scholar] [CrossRef]
  17. Bakshi, S.; Banik, C.; Rathke, S.J.; Laird, D.A. Arsenic Sorption on Zero-Valent Iron-Biochar Complexes. Water Res. 2018, 137, 153–163. [Google Scholar] [CrossRef]
  18. Nzediegwu, C.; Naeth, M.A.; Chang, S.X. Lead(II) Adsorption on Microwave-Pyrolyzed Biochars and Hydrochars Depends on Feedstock Type and Production Temperature. J. Hazard. Mater. 2021, 412, 125255. [Google Scholar] [CrossRef]
  19. Xiao, Y.; Xue, Y.; Gao, F.; Mosa, A. Sorption of Heavy Metal Ions onto Crayfish Shell Biochar: Effect of Pyrolysis Temperature, PH and Ionic Strength. J. Taiwan Inst. Chem. Eng. 2017, 80, 114–121. [Google Scholar] [CrossRef]
  20. Yang, Q.; Mašek, O.; Zhao, L.; Nan, H.; Yu, S.; Yin, J.; Li, Z.; Cao, X. Country-Level Potential of Carbon Sequestration and Environmental Benefits by Utilizing Crop Residues for Biochar Implementation. Appl. Energy 2021, 282, 116275. [Google Scholar] [CrossRef]
  21. Ahmad, M.; Rajapaksha, A.U.; Lim, J.E.; Zhang, M.; Bolan, N.; Mohan, D.; Vithanage, M.; Lee, S.S.; Ok, Y.S. Biochar as a Sorbent for Contaminant Management in Soil and Water: A Review. Chemosphere 2014, 99, 19–33. [Google Scholar] [CrossRef] [PubMed]
  22. 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]
  23. Shen, Z.; Zhang, J.; Hou, D.; Tsang, D.C.W.; Ok, Y.S.; Alessi, D.S. Synthesis of MgO-Coated Corncob Biochar and Its Application in Lead Stabilization in a Soil Washing Residue. Environ. Int. 2019, 122, 357–362. [Google Scholar] [CrossRef] [PubMed]
  24. Zhou, N.; Chen, H.; Xi, J.; Yao, D.; Zhou, Z.; Tian, Y.; Lu, X. Biochars with Excellent Pb(II) Adsorption Property Produced from Fresh and Dehydrated Banana Peels via Hydrothermal Carbonization. Bioresour. Technol. 2017, 232, 204–210. [Google Scholar] [CrossRef] [PubMed]
  25. Liu, P.; Rao, D.; Zou, L.; Teng, Y.; Yu, H. Capacity and Potential Mechanisms of Cd(II) Adsorption from Aqueous Solution by Blue Algae-Derived Biochars. Sci. Total Environ. 2021, 767, 145447. [Google Scholar] [CrossRef]
  26. Mokrzycki, J.; Michalak, I.; Rutkowski, P. Biochars Obtained from Freshwater Biomass—Green Macroalga and Hornwort as Cr(III) Ions Sorbents. Biomass Convers. Bioref. 2021, 11, 301–313. [Google Scholar] [CrossRef]
  27. Michalak, I.; Baśladyńska, S.; Mokrzycki, J.; Rutkowski, P. Biochar from A Freshwater Macroalga as A Potential Biosorbent for Wastewater Treatment. Water 2019, 11, 1390. [Google Scholar] [CrossRef]
  28. Mokrzycki, J.; Michalak, I.; Rutkowski, P. Tomato Green Waste Biochars as Sustainable Trivalent Chromium Sorbents. Environ. Sci. Pollut. Res. 2021, 28, 24245–24255. [Google Scholar] [CrossRef]
  29. Rajapaksha, A.U.; Alam, M.S.; Chen, N.; Alessi, D.S.; Igalavithana, A.D.; Tsang, D.C.W.; Ok, Y.S. Removal of Hexavalent Chromium in Aqueous Solutions Using Biochar: Chemical and Spectroscopic Investigations. Sci. Total Environ. 2018, 625, 1567–1573. [Google Scholar] [CrossRef]
  30. Wang, S.; Gao, B.; Zimmerman, A.R.; Li, Y.; Ma, L.; Harris, W.G.; Migliaccio, K.W. Removal of Arsenic by Magnetic Biochar Prepared from Pinewood and Natural Hematite. Bioresour. Technol. 2015, 175, 391–395. [Google Scholar] [CrossRef]
  31. Zhou, M.; Zhang, C.; Yuan, Y.; Mao, X.; Li, Y.; Wang, N.; Wang, S.; Wang, X. Pinewood Outperformed Bamboo as Feedstock to Prepare Biochar-Supported Zero-Valent Iron for Cr6+ Reduction. Environ. Res. 2020, 187, 109695. [Google Scholar] [CrossRef] [PubMed]
  32. Deng, R.; Huang, D.; Wan, J.; Xue, W.; Wen, X.; Liu, X.; Chen, S.; Lei, L.; Zhang, Q. Recent Advances of Biochar Materials for Typical Potentially Toxic Elements Management in Aquatic Environments: A Review. J. Clean. Prod. 2020, 255, 119523. [Google Scholar] [CrossRef]
  33. Zhang, X.; Li, Y.; Wu, M.; Pang, Y.; Hao, Z.; Hu, M.; Qiu, R.; Chen, Z. Enhanced Adsorption of Tetracycline by an Iron and Manganese Oxides Loaded Biochar: Kinetics, Mechanism and Column Adsorption. Bioresour. Technol. 2021, 320, 124264. [Google Scholar] [CrossRef]
  34. Li, L.; Cao, G.; Zhu, R. Adsorption of Cr(VI) from Aqueous Solution by a Litchi Shell-Based Adsorbent. Environ. Res. 2021, 196, 110356. [Google Scholar] [CrossRef]
  35. Liu, X.; Gao, M.; Qiu, W.; Khan, Z.H.; Liu, N.; Lin, L.; Song, Z. Fe-Mn-Ce Oxide-Modified Biochar Composites as Efficient Adsorbents for Removing As(III) from Water: Adsorption Performance and Mechanisms. Environ. Sci. Pollut. Res. 2019, 26, 17373–17382. [Google Scholar] [CrossRef] [PubMed]
  36. Zhu, Y.; Dai, W.; Deng, K.; Pan, T.; Guan, Z. Efficient Removal of Cr(VI) from Aqueous Solution by Fe-Mn Oxide-Modified Biochar. Water Air Soil Pollut. 2020, 231, 61. [Google Scholar] [CrossRef]
  37. Thotagamuge, R.; Kooh, M.R.R.; Mahadi, A.H.; Lim, C.M.; Abu, M.; Jan, A.; Hanipah, A.H.A.; Khiong, Y.Y.; Shofry, A. Copper Modified Activated Bamboo Charcoal to Enhance Adsorption of Heavy Metals from Industrial Wastewater. Environ. Nanotechnol. Monit. Manag. 2021, 16, 100562. [Google Scholar] [CrossRef]
  38. Masanizan, A.; Lim, C.M.; Kooh, M.R.R.; Mahadi, A.H.; Thotagamuge, R. The Removal of Ruthenium-Based Complexes N3 Dye from DSSC Wastewater Using Copper Impregnated KOH-Activated Bamboo Charcoal. Water Air Soil Pollut. 2021, 232, 388. [Google Scholar] [CrossRef]
  39. Tan, X.; Liu, Y.; Gu, Y.; Xu, Y.; Zeng, G.; Hu, X.; Liu, S.; Wang, X.; Liu, S.; Li, J. Biochar-Based Nano-Composites for the Decontamination of Wastewater: A Review. Bioresour. Technol. 2016, 212, 318–333. [Google Scholar] [CrossRef]
  40. Younis, S.A.; Kim, K.-H.; Shaheen, S.M.; Antoniadis, V.; Tsang, Y.F.; Rinklebe, J.; Deep, A.; Brown, R.J.C. Advancements of Nanotechnologies in Crop Promotion and Soil Fertility: Benefits, Life Cycle Assessment, and Legislation Policies. Renew. Sustain. Energy Rev. 2021, 152, 111686. [Google Scholar] [CrossRef]
  41. Wang, H.; Gao, B.; Wang, S.; Fang, J.; Xue, Y.; Yang, K. Removal of Pb(II), Cu(II), and Cd(II) from Aqueous Solutions by Biochar Derived from KMnO4 Treated Hickory Wood. Bioresour. Technol. 2015, 197, 356–362. [Google Scholar] [CrossRef] [PubMed]
  42. Zhu, Y.; Fan, W.; Zhang, K.; Xiang, H.; Wang, X. Nano-Manganese Oxides-Modified Biochar for Efficient Chelated Copper Citrate Removal from Water by Oxidation-Assisted Adsorption Process. Sci. Total Environ. 2020, 709, 136154. [Google Scholar] [CrossRef] [PubMed]
  43. Wan, S.; Qiu, L.; Tang, G.; Chen, W.; Li, Y.; Gao, B.; He, F. Ultrafast Sequestration of Cadmium and Lead from Water by Manganese Oxide Supported on a Macro-Mesoporous Biochar. Chem. Eng. J. 2020, 387, 124095. [Google Scholar] [CrossRef]
  44. Shi, S.; Yang, J.; Liang, S.; Li, M.; Gan, Q.; Xiao, K.; Hu, J. Enhanced Cr(VI) Removal from Acidic Solutions Using Biochar Modified by Fe3O4@SiO2-NH2 Particles. Sci. Total Environ. 2018, 628–629, 499–508. [Google Scholar] [CrossRef]
  45. Fan, Z.; Zhang, Q.; Gao, B.; Li, M.; Liu, C.; Qiu, Y. Removal of Hexavalent Chromium by Biochar Supported nZVI Composite: Batch and Fixed-Bed Column Evaluations, Mechanisms, and Secondary Contamination Prevention. Chemosphere 2019, 217, 85–94. [Google Scholar] [CrossRef]
  46. Sinha, R.; Kumar, R.; Abhishek, K.; Shang, J.; Bhattacharya, S.; Sengupta, S.; Kumar, N.; Singh, R.K.; Mallick, J.; Kar, M.; et al. Single-Step Synthesis of Activated Magnetic Biochar Derived from Rice Husk for Hexavalent Chromium Adsorption: Equilibrium Mechanism, Kinetics, and Thermodynamics Analysis. Groundw. Sustain. Dev. 2022, 18, 100796. [Google Scholar] [CrossRef]
  47. Li, C.; Zhang, L.; Gao, Y.; Li, A. Facile Synthesis of Nano ZnO/ZnS Modified Biochar by Directly Pyrolyzing of Zinc Contaminated Corn Stover for Pb(II), Cu(II) and Cr(VI) Removals. Waste Manag. 2018, 79, 625–637. [Google Scholar] [CrossRef]
  48. Zhu, N.; Yan, T.; Qiao, J.; Cao, H. Adsorption of Arsenic, Phosphorus and Chromium by Bismuth Impregnated Biochar: Adsorption Mechanism and Depleted Adsorbent Utilization. Chemosphere 2016, 164, 32–40. [Google Scholar] [CrossRef]
  49. Guo, N.; Lv, X.; Yang, Q.; Xu, X.; Song, H. Effective Removal of Hexavalent Chromium from Aqueous Solution by ZnCl2 Modified Biochar: Effects and Response Sequence of the Functional Groups. J. Mol. Liq. 2021, 334, 116149. [Google Scholar] [CrossRef]
  50. Wang, X.; Xu, J.; Liu, J.; Liu, J.; Xia, F.; Wang, C.; Dahlgren, R.A.; Liu, W. Mechanism of Cr(VI) Removal by Magnetic Greigite/Biochar Composites. Sci. Total Environ. 2020, 700, 134414. [Google Scholar] [CrossRef]
  51. Hoang, L.P.; Van, H.T.; Nguyen, L.H.; Mac, D.-H.; Vu, T.T.; Ha, L.T.; Nguyen, X.C. Removal of Cr(VI) from Aqueous Solution Using Magnetic Modified Biochar Derived from Raw Corncob. New J. Chem. 2019, 43, 18663–18672. [Google Scholar] [CrossRef]
  52. Zhang, X.; Lv, L.; Qin, Y.; Xu, M.; Jia, X.; Chen, Z. Removal of Aqueous Cr(VI) by a Magnetic Biochar Derived from Melia Azedarach Wood. Bioresour. Technol. 2018, 256, 1–10. [Google Scholar] [CrossRef] [PubMed]
  53. Yang, Y.; Chen, N.; Feng, C.; Li, M.; Gao, Y. Chromium Removal Using a Magnetic Corncob Biochar/Polypyrrole Composite by Adsorption Combined with Reduction: Reaction Pathway and Contribution Degree. Colloids Surf. A Physicochem. Eng. Asp. 2018, 556, 201–209. [Google Scholar] [CrossRef]
  54. Dong, F.-X.; Yan, L.; Zhou, X.-H.; Huang, S.-T.; Liang, J.-Y.; Zhang, W.-X.; Guo, Z.-W.; Guo, P.-R.; Qian, W.; Kong, L.-J.; et al. Simultaneous Adsorption of Cr(VI) and Phenol by Biochar-Based Iron Oxide Composites in Water: Performance, Kinetics and Mechanism. J. Hazard. Mater. 2021, 416, 125930. [Google Scholar] [CrossRef] [PubMed]
  55. Cuong, D.V.; Wu, P.-C.; Chen, L.-I.; Hou, C.-H. Active MnO2/Biochar Composite for Efficient As(III) Removal: Insight into the Mechanisms of Redox Transformation and Adsorption. Water Res. 2021, 188, 116495. [Google Scholar] [CrossRef] [PubMed]
  56. Li, W.; Liu, J.; Zhao, D. Mesoporous Materials for Energy Conversion and Storage Devices. Nat. Rev. Mater. 2016, 1, 16023. [Google Scholar] [CrossRef]
  57. Trubetskaya, A.; Jensen, P.A.; Jensen, A.D.; Glarborg, P.; Larsen, F.H.; Andersen, M.L. Characterization of Free Radicals by Electron Spin Resonance Spectroscopy in Biochars from Pyrolysis at High Heating Rates and at High Temperatures. Biomass Bioenergy 2016, 94, 117–129. [Google Scholar] [CrossRef]
  58. Klüpfel, L.; Keiluweit, M.; Kleber, M.; Sander, M. Redox Properties of Plant Biomass-Derived Black Carbon (Biochar). Environ. Sci. Technol. 2014, 48, 5601–5611. [Google Scholar] [CrossRef] [PubMed]
  59. Kamran, U.; Park, S.-J. MnO2-Decorated Biochar Composites of Coconut Shell and Rice Husk: An Efficient Lithium Ions Adsorption-Desorption Performance in Aqueous Media. Chemosphere 2020, 260, 127500. [Google Scholar] [CrossRef]
  60. Mohan, D.; Abhishek, K.; Sarswat, A.; Patel, M.; Singh, P.; Pittman, C.U. Biochar Production and Applications in Soil Fertility and Carbon Sequestration-a Sustainable Solution to Crop-Residue Burning in India. RSC Adv. 2018, 8, 508–520. [Google Scholar] [CrossRef]
  61. Niazi, N.K.; Bibi, I.; Shahid, M.; Ok, Y.S.; Shaheen, S.M.; Rinklebe, J.; Wang, H.; Murtaza, B.; Islam, E.; Farrakh Nawaz, M.; et al. Arsenic Removal by Japanese Oak Wood Biochar in Aqueous Solutions and Well Water: Investigating Arsenic Fate Using Integrated Spectroscopic and Microscopic Techniques. Sci. Total Environ. 2018, 621, 1642–1651. [Google Scholar] [CrossRef] [PubMed]
  62. Zhong, D.; Jiang, Y.; Zhao, Z.; Wang, L.; Chen, J.; Ren, S.; Liu, Z.; Zhang, Y.; Tsang, D.C.W.; Crittenden, J.C. PH Dependence of Arsenic Oxidation by Rice-Husk-Derived Biochar: Roles of Redox-Active Moieties. Environ. Sci. Technol. 2019, 53, 9034–9044. [Google Scholar] [CrossRef] [PubMed]
  63. Liu, Z.; Zhong, X.; Wang, Y.; Ding, Z.; Wang, C.; Wang, G.; Liao, S. An Efficient Adsorption of Manganese Oxides/Activated Carbon Composite for Lead(II) Ions from Aqueous Solution. Arab. J. Sci. Eng. 2018, 43, 2155–2165. [Google Scholar] [CrossRef]
  64. Duan, S.; Ma, W.; Pan, Y.; Meng, F.; Yu, S.; Wu, L. Synthesis of Magnetic Biochar from Iron Sludge for the Enhancement of Cr (VI) Removal from Solution. J. Taiwan Inst. Chem. Eng. 2017, 80, 835–841. [Google Scholar] [CrossRef]
  65. Shaheen, S.M.; Natasha; Mosa, A.; El-Naggar, A.; Faysal Hossain, M.; Abdelrahman, H.; Khan Niazi, N.; Shahid, M.; Zhang, T.; Fai Tsang, Y.; et al. Manganese Oxide-Modified Biochar: Production, Characterization and Applications for the Removal of Pollutants from Aqueous Environments—A Review. Bioresour. Technol. 2022, 346, 126581. [Google Scholar] [CrossRef]
  66. Miretzky, P.; Cirelli, A.F. Cr(VI) and Cr(III) Removal from Aqueous Solution by Raw and Modified Lignocellulosic Materials: A Review. J. Hazard. Mater. 2010, 180, 1–19. [Google Scholar] [CrossRef]
  67. Wang, X.S.; Chen, L.F.; Li, F.Y.; Chen, K.L.; Wan, W.Y.; Tang, Y.J. Removal of Cr (VI) with Wheat-Residue Derived Black Carbon: Reaction Mechanism and Adsorption Performance. J. Hazard. Mater. 2010, 175, 816–822. [Google Scholar] [CrossRef]
  68. Pehlivan, E.; Kahraman, H.; Pehlivan, E. Sorption Equilibrium of Cr(VI) Ions on Oak Wood Charcoal (Carbo ligni) and Charcoal Ash as Low-Cost Adsorbents. Fuel Process. Technol. 2011, 92, 65–70. [Google Scholar] [CrossRef]
  69. Xiao, J.; Hu, R.; Chen, G.; Xing, B. Facile Synthesis of Multifunctional Bone Biochar Composites Decorated with Fe/Mn Oxide Micro-Nanoparticles: Physicochemical Properties, Heavy Metals Sorption Behavior and Mechanism. J. Hazard. Mater. 2020, 399, 123067. [Google Scholar] [CrossRef]
  70. Xu, R.; Xiao, S.; Yuan, J.; Zhao, A. Adsorption of Methyl Violet from Aqueous Solutions by the Biochars Derived from Crop Residues. Bioresour. Technol. 2011, 102, 10293–10298. [Google Scholar] [CrossRef]
  71. Park, C.M.; Han, J.; Chu, K.H.; Al-Hamadani, Y.A.J.; Her, N.; Heo, J.; Yoon, Y. Influence of Solution PH, Ionic Strength, and Humic Acid on Cadmium Adsorption onto Activated Biochar: Experiment and Modeling. J. Ind. Eng. Chem. 2017, 48, 186–193. [Google Scholar] [CrossRef]
  72. Gheju, M.; Balcu, I.; Mosoarca, G. Removal of Cr(VI) from Aqueous Solutions by Adsorption on MnO2. J. Hazard. Mater. 2016, 310, 270–277. [Google Scholar] [CrossRef] [PubMed]
  73. Zhao, J.; Boada, R.; Cibin, G.; Palet, C. Enhancement of Selective Adsorption of Cr Species via Modification of Pine Biomass. Sci. Total Environ. 2021, 756, 143816. [Google Scholar] [CrossRef] [PubMed]
  74. Cheng, D.; Ngo, H.H.; Guo, W.; Chang, S.W.; Nguyen, D.D.; Zhang, X.; Varjani, S.; Liu, Y. Feasibility Study on a New Pomelo Peel Derived Biochar for Tetracycline Antibiotics Removal in Swine Wastewater. Sci. Total Environ. 2020, 720, 137662. [Google Scholar] [CrossRef]
  75. Jin, H.; Capareda, S.; Chang, Z.; Gao, J.; Xu, Y.; Zhang, J. Biochar Pyrolytically Produced from Municipal Solid Wastes for Aqueous As(V) Removal: Adsorption Property and Its Improvement with KOH Activation. Bioresour. Technol. 2014, 169, 622–629. [Google Scholar] [CrossRef] [PubMed]
  76. Xi, J.; He, M.; Lin, C. Adsorption of Antimony(III) and Antimony(V) on Bentonite: Kinetics, Thermodynamics and Anion Competition. Microchem. J. 2011, 97, 85–91. [Google Scholar] [CrossRef]
  77. Yoon, K.; Cho, D.-W.; Tsang, D.C.W.; Bolan, N.; Rinklebe, J.; Song, H. Fabrication of Engineered Biochar from Paper Mill Sludge and Its Application into Removal of Arsenic and Cadmium in Acidic Water. Bioresour. Technol. 2017, 246, 69–75. [Google Scholar] [CrossRef]
  78. Shen, C.; Chen, H.; Wu, S.; Wen, Y.; Li, L.; Jiang, Z.; Li, M.; Liu, W. Highly Efficient Detoxification of Cr(VI) by Chitosan–Fe(III) Complex: Process and Mechanism Studies. J. Hazard. Mater. 2013, 244–245, 689–697. [Google Scholar] [CrossRef]
  79. Chen, N.; Cao, S.; Zhang, L.; Peng, X.; Wang, X.; Ai, Z.; Zhang, L. Structural Dependent Cr(VI) Adsorption and Reduction of Biochar: Hydrochar versus Pyrochar. Sci. Total Environ. 2021, 783, 147084. [Google Scholar] [CrossRef]
  80. Jia, X.; Zhou, J.; Liu, J.; Liu, P.; Yu, L.; Wen, B.; Feng, Y. The Antimony Sorption and Transport Mechanisms in Removal Experiment by Mn-Coated Biochar. Sci. Total Environ. 2020, 724, 138158. [Google Scholar] [CrossRef]
  81. Liu, P.; Liu, J.; Zhang, B.; Zong, W.; Xu, S.; Liu, Y.; Cao, S. Enhanced Electroluminescent Performance by Doping Organic Conjugated Ionic Compound into Graphene Oxide Hole-Injecting Layer. J. Mater. Sci. 2019, 54, 12688–12697. [Google Scholar] [CrossRef]
  82. Song, Z.; Lian, F.; Yu, Z.; Zhu, L.; Xing, B.; Qiu, W. Synthesis and Characterization of a Novel MnOx-Loaded Biochar and Its Adsorption Properties for Cu2+ in Aqueous Solution. Chem. Eng. J. 2014, 242, 36–42. [Google Scholar] [CrossRef]
  83. Zhang, S.; Li, X.; Chen, J.P. An XPS Study for Mechanisms of Arsenate Adsorption onto a Magnetite-Doped Activated Carbon Fiber. J. Colloid Interface Sci. 2010, 343, 232–238. [Google Scholar] [CrossRef] [PubMed]
  84. Lee, H.-S.; Shin, H.-S. Competitive Adsorption of Heavy Metals onto Modified Biochars: Comparison of Biochar Properties and Modification Methods. J. Environ. Manag. 2021, 299, 113651. [Google Scholar] [CrossRef]
  85. Choi, K.; Yong Lee, S.; Kim, H.; Bong Lee, K.; Choi, J.-W.; Jung, K.-W. Mechanistic Insights into the Simultaneous Removal of As(V) and Cr(VI) Oxyanions by a Novel Hierarchical Corolla-like MnO2-Decorated Porous Magnetic Biochar Composite: A Combined Experimental and Density Functional Theory Study. Appl. Surf. Sci. 2022, 578, 151991. [Google Scholar] [CrossRef]
  86. Zhou, L.; Liu, Y.; Liu, S.; Yin, Y.; Zeng, G.; Tan, X.; Hu, X.; Hu, X.; Jiang, L.; Ding, Y.; et al. Investigation of the Adsorption-Reduction Mechanisms of Hexavalent Chromium by Ramie Biochars of Different Pyrolytic Temperatures. Bioresour. Technol. 2016, 218, 351–359. [Google Scholar] [CrossRef]
  87. Yu, Z.; Zhou, L.; Huang, Y.; Song, Z.; Qiu, W. Effects of a Manganese Oxide-Modified Biochar Composite on Adsorption of Arsenic in Red Soil. J. Environ. Manag. 2015, 163, 155–162. [Google Scholar] [CrossRef]
  88. Ding, Z.; Sun, G.; Fu, F.; Ye, C. Phase Transformation of Cr(VI)-Adsorbed Ferr Ihydr Ite in the Presence of Mn(II): Fate of Mn(II) and Cr(VI). J. Environ. Sci. 2022, 113, 251–259. [Google Scholar] [CrossRef]
Figure 1. The scheme of the process to synthesis of MnOx-modified biochar.
Figure 1. The scheme of the process to synthesis of MnOx-modified biochar.
Water 14 02507 g001
Figure 2. SEM images of biochars: (a) BC, (b) MnOBC-1, and (c) MnOBC-2.
Figure 2. SEM images of biochars: (a) BC, (b) MnOBC-1, and (c) MnOBC-2.
Water 14 02507 g002
Figure 3. FT-IR (a) and X-ray diffraction (XRD) (b) of the samples.
Figure 3. FT-IR (a) and X-ray diffraction (XRD) (b) of the samples.
Water 14 02507 g003
Figure 4. Effect of pH (a) and ionic strength (b) on Cr(VI) sorption by BC, MnOBC-1, and MnOBC-2. Conditions: C0(Cr(VI)) = 50 mg L−1, sorbent dose = 1 g L−1, and temperature = 298 K.
Figure 4. Effect of pH (a) and ionic strength (b) on Cr(VI) sorption by BC, MnOBC-1, and MnOBC-2. Conditions: C0(Cr(VI)) = 50 mg L−1, sorbent dose = 1 g L−1, and temperature = 298 K.
Water 14 02507 g004
Figure 5. Sorption kinetics of Cr(VI) on BC, MnOBC-1, and MnOBC-2. Conditions: C0(Cr(VI)) = 50 mg L−1, sorbent dose = 1 g L−1, pH = 2.0 ± 0.2, and temperature = 298 K.
Figure 5. Sorption kinetics of Cr(VI) on BC, MnOBC-1, and MnOBC-2. Conditions: C0(Cr(VI)) = 50 mg L−1, sorbent dose = 1 g L−1, pH = 2.0 ± 0.2, and temperature = 298 K.
Water 14 02507 g005
Figure 6. Sorption isotherms of Cr(VI) on BC, MnOBC-1, and MnOBC-2. Conditions: C0(Cr(VI)) = 0–200 mg L−1, sorbent dose = 1 g L−1, pH = 2.0 ± 0.2, and temperature = 298 K.
Figure 6. Sorption isotherms of Cr(VI) on BC, MnOBC-1, and MnOBC-2. Conditions: C0(Cr(VI)) = 0–200 mg L−1, sorbent dose = 1 g L−1, pH = 2.0 ± 0.2, and temperature = 298 K.
Water 14 02507 g006
Figure 7. Cr2p3/2 spectra after Cr(VI) sorption: (a) BC, (b) MnOBC-1, and (c) MnOBC-2.
Figure 7. Cr2p3/2 spectra after Cr(VI) sorption: (a) BC, (b) MnOBC-1, and (c) MnOBC-2.
Water 14 02507 g007
Figure 8. (a) Full-range of XPS spectra, (b) Mn2p3/2 spectra before and after Cr(VI) sorption by MnOBC-2, (c) C1s spectra before and after Cr(VI) sorption by MnOBC-2, and (d) O1s spectra before and after Cr(VI) sorption by MnOBC-2.
Figure 8. (a) Full-range of XPS spectra, (b) Mn2p3/2 spectra before and after Cr(VI) sorption by MnOBC-2, (c) C1s spectra before and after Cr(VI) sorption by MnOBC-2, and (d) O1s spectra before and after Cr(VI) sorption by MnOBC-2.
Water 14 02507 g008
Figure 9. TEM-EDS of of BC after sorption of Cr(VI) (a), MnOBC-1 after sorption of Cr(VI) (b), and MnOBC-2 after sorption of Cr(VI) (c).
Figure 9. TEM-EDS of of BC after sorption of Cr(VI) (a), MnOBC-1 after sorption of Cr(VI) (b), and MnOBC-2 after sorption of Cr(VI) (c).
Water 14 02507 g009
Table 1. Study on Cr(VI) removal by different biochar.
Table 1. Study on Cr(VI) removal by different biochar.
Biochar
Feedstock
Main Chemicals UsedReported Mechanismsqmax (mg g−1)Reference
Phoenix tree leavesFeCl3·6H2Oelectrostatic attraction, reduction, and chelation27.2[44]
Sewage sludgeFeCl3·6H2Oelectrostatic attraction, complexation, and reduction reaction31.53[45]
Rice huskFeCl3·6H2O
ZnCl2
electrostatic attraction and ion exchange9.97[46]
Corn stoverZnSO4electrostatic attraction and complexation24.5[47]
Wheat strawBi2O3electrostatic attraction and reduction reaction12.23[48]
Sea buckthorn stonesZnCl2electrostatic attraction, complexation, and reduction reaction19.3[49]
Rice huskFeSO4·7H2Oelectrostatic attraction and reduction reaction23.25[50]
CorncobFeCl3·6H2Oelectrostatic attraction, ion exchange and adsorption coupled-reduction25.94[51]
Melia azedarach woodFe(NO3)3·9H2Oreduction reaction25.27[52]
CorncobsFe(NO3)3, FeCl3·6H2O and pyrroleion exchange, chelation, complexation, and reduction reaction19.23[53]
Rice huskMn(NO3)2electrostatic attraction, complexation, and reduction reaction28.58This study
Table 2. Pore structures of BC and MnOBCs.
Table 2. Pore structures of BC and MnOBCs.
BCMnOBC-1MnOBC-2
BET specific area (m2 g−1)33.67302.02299.56
Total pore volume (cm3 g−1)0.0250.1300.134
Average pore diameter (nm)3.021.721.79
Table 3. Regressed kinetic parameters for Cr(VI) sorption onto BC and MnOBCs.
Table 3. Regressed kinetic parameters for Cr(VI) sorption onto BC and MnOBCs.
The Pseudo-First-Order ModelThe Pseudo-Second-Order ModelElovich Model
qe (mg g−1)k1R2qe (mg g−1)k2R2αβR2
BC5.120.140.85705.370.0370.9334182.832.300.9799
MnOBC-113.840.0180.750315.200.00160.85942.140.470.9802
MnOBC-222.040.0200.728123.760.00130.84295.020.310.9821
Table 4. Langmuir and Freundlich equation parameters calculated using nonlinear curve-fitting for sorption isotherms for BC and MnOBCs.
Table 4. Langmuir and Freundlich equation parameters calculated using nonlinear curve-fitting for sorption isotherms for BC and MnOBCs.
LangmuirFreundlich
qmax (mg g−1)KLR2KF1/nR2
BC10.100.0820.85822.710.230.9821
MnOBC-120.690.570.907910.250.170.9886
MnOBC-228.580.300.886311.540.210.9868
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Fan, J.; Qin, L.; Duan, T.; Qi, Z.; Zou, L. Preparation of MnOx-Modified Biochar and Its Removal Mechanism for Cr(VI) in Aqueous Solution. Water 2022, 14, 2507. https://doi.org/10.3390/w14162507

AMA Style

Fan J, Qin L, Duan T, Qi Z, Zou L. Preparation of MnOx-Modified Biochar and Its Removal Mechanism for Cr(VI) in Aqueous Solution. Water. 2022; 14(16):2507. https://doi.org/10.3390/w14162507

Chicago/Turabian Style

Fan, Jianxin, Liang Qin, Ting Duan, Zenglin Qi, and Lan Zou. 2022. "Preparation of MnOx-Modified Biochar and Its Removal Mechanism for Cr(VI) in Aqueous Solution" Water 14, no. 16: 2507. https://doi.org/10.3390/w14162507

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