Next Article in Journal
Molecular Insights and Clinical Outcomes of Drugs of Abuse Adulteration: New Trends and New Psychoactive Substances
Next Article in Special Issue
Photocatalytic Performance of Undoped and Al-Doped ZnO Nanoparticles in the Degradation of Rhodamine B under UV-Visible Light:The Role of Defects and Morphology
Previous Article in Journal
Osteoblast Demineralization Induced by Oxidized High-Density Lipoprotein via the Inflammatory Pathway Is Suppressed by Adiponectin
Previous Article in Special Issue
The Synergistic Effect of Adsorption-Photocatalysis for Removal of Organic Pollutants on Mesoporous Cu2V2O7/Cu3V2O8/g-C3N4 Heterojunction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

One-Step Synthesis of Nitrogen-Doped Porous Biochar Based on N-Doping Co-Activation Method and Its Application in Water Pollutants Control

1
College of Life Sciences, Jilin Agricultural University, Changchun 130118, China
2
Key Laboratory of Straw Comprehensive Utilization and Black Soil Conservation, Ministry of Education, Jilin Agricultural University, Changchun 130118, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(23), 14618; https://doi.org/10.3390/ijms232314618
Submission received: 18 October 2022 / Revised: 11 November 2022 / Accepted: 17 November 2022 / Published: 23 November 2022

Abstract

:
In this work, birch bark (BB) was used for the first time to prepare porous biochars via different one-step methods including direct activation (BBB) and N-doping co-activation (N-BBB). The specific surface area and total pore volume of BBB and N-BBB were 2502.3 and 2292.7 m2/g, and 1.1389 and 1.0356 cm3/g, respectively. When removing synthetic methyl orange (MO) dye and heavy metal Cr6+, both BBB and N-BBB showed excellent treatment ability. The maximum adsorption capacities of BBB and N-BBB were 836.9 and 858.3 mg/g for MO, and 141.1 and 169.1 mg/g for Cr6+, respectively, which were higher than most previously reported biochar adsorbents. The probable adsorption mechanisms, including pore filling, π–π interaction, H-bond interaction, and electrostatic attraction, supported the biochars’ demonstrated high performance. In addition, after five recycles, the removal rates remained above 80%, which showed the high stability of the biochars. This work verified the feasibility of the one-step N-doping co-activation method to prepare high-performance biochars, and two kinds of biochars with excellent performance (BBB and N-BBB) were prepared. More importantly, this method provides new directions and ideas for the development and utilization of other biomasses.

1. Introduction

Water pollution is a serious threat to the ecosystem and human safety [1]. Organic pollutants (aromatic species) and heavy metals in water have become the focus of global attention due to their strong teratogenicity, carcinogenicity, and non-degradability [2,3]. Cr6+ is one of the main forms of Cr. Because of its high solubility, easy accumulation, and difficulty degrading in the ecosystem, Cr6+ is one of the main toxic heavy metals that pollute water bodies [4]. Cr6+ ions are discharged into the water in large quantities during industrial processes such as metallurgy and electroplating [5]. Similarly, synthetic dyes are widely used in the printing, dyeing, and textile industries [6]. Many synthetic dyes have caused irreversible threats to human safety, such as skin allergy and cancer, and have become a research hotspot in the field of water pollution remediation [7,8]. Therefore, it is important to remove Cr6+ and synthetic dyes to purify water.
At present, the removal methods for Cr6+ and synthetic dyes from polluted water mainly include biological treatment, chemical treatment, and physical treatment [9,10,11]. As one of the physical treatment methods, the adsorption method is widely used in water pollution treatment because of its low cost, lack of by-products, and mild conditions [12,13]. It is also considered an ideal method to effectively remove Cr6+ and synthetic dyes. Many materials, such as graphene oxide and carbon nanotubes, are used to treat water pollution due to their excellent adsorption properties on pollutants, but the high cost often limits their application [14,15]. In contrast, biochar is attracting increasing attention as an excellent adsorbent with a large specific surface area, abundant functional groups, and a well-developed pore structure [16,17]. Unfortunately, due to the limited adsorption sites, the adsorption performance is often limited. Appropriately modifying and improving the adsorption efficiency of biochar is the key to determining whether it can become a widely used adsorbent.
As an effective strategy to improve biochar performance, nitrogen doping (N-doping) technology has been applied in environmental remediation, such as adsorption, catalysis, and electrochemistry [18,19]. N hetero-atoms provide more binding sites on the biochar surface, which enhances the hydrophilicity and adsorption capacity of the compound by forming a more stable complex with pollutants [20]. At the same time, the N-doping method can adjust part of the electronic structure in the carbon skeleton, and further improve the reactivity of the biochar [21]. The pore structure, specific surface area, and other physical and chemical properties of biochars are also crucial factors that not only affect the adsorption capacity but also determine the degree of adsorbed pollutant diffusion [22,23]. The pore-forming effect of common activators (such as NaOH and KOH) in biochar preparation is beyond doubt [24,25]. However, there are few studies on biochar prepared with one-step pyrolysis and N-doping.
To remedy this, we explored a novel biochar preparation strategy to improve the performance of biochars, that is, N-doping co-activation. N-doping co-activation refers to the incorporation of nitrogen sources in the activation process in order to complete nitrogen doping and activation at the same time and successfully prepare nitrogen-doped porous biochars in one step. In this study, Birch bark was used as the carbon source for biochar preparation for the first time. N-doped porous biochar was synthesized with one-step co-activation pyrolysis by adding an activator (NaOH) and a nitrogen source (urea) in the preparation process. The focus of this study is to determine the feasibility of one-step N-doping co-activation in biochar preparation and its superiority in water pollution treatment, in order to provide a more effective way to deal with water pollution.

2. Results and Discussion

2.1. Preparation of CBB, BBB, and N-BBB

The sample preparation process is shown in Figure 1 and can be divided into three processes: direct carbonization, activation, and N-doping co-activation. The probable chemical reactions during the processes are summarized as follows:
Activation process [26,27,28]: In the high-temperature pyrolysis process, NaOH will generate Na2O and water, and then be further ionized to form Na+ and OH. These ions will migrate or insert carbon precursors and react with them to generate carbon dioxide and water. The carbon dioxide further reacts with the oxide produced by NaOH (Na2O) to form carbonate (Na2CO3), which etches the carbonized sample. Finally, microporous, or mesoporous, structures are formed in the carbon.
N-doping process [29,30,31]: The decomposition of CO(NH2)2 at a high temperature produces solid and gaseous substances. Through the diffusion of the gas, the surface of the biochar develops various wrinkles and pores. The NH3 released during the pyrolysis process combines with the -OH group in lignocellulose (the main component of BB) to form C-NH2, and further generate C=N-C [20]. Eventually, these nitrogenous groups form various N-bond configurations including pyrrolic-N, pyridinic-N, and graphite-N.

2.2. Results of Characterizations

The microscopic morphology and elemental composition of the samples were observed by SEM and EDS, as shown in Figure 2. BB had a rough, irregular surface and uneven lamellar structure. After carbonization (Figure 2B), CBB began to show a dense and smooth surface morphology. This can be explained by the fact that the main composition of BB is lignocellulose, consisting mainly of cellulose composed of glucose and hemicellulose composed of xylose [32]. During the high-temperature pyrolysis process, denatured glucose and xylose changed the morphology of CBB [33]. With the activation process treatment, especially in the presence of NaOH, BBB, and N-BBB (Figure 2C,E), many cracks and fragmentation occurred on the surface due to a severe reaction phenomenon. BBB and N-BBB were mainly composed of C, O, and N elements. However, because of different treatments, their nitrogen content was different. The C, O, and N contents of BBB and N-BBB were 84.62%, 13.52%, and 1.87%, and 84.16%, 12.14%, and 3.70%, respectively. Compared with BBB, the content of N in N-BBB nearly doubled. The above results not only proved the success of the activation process but also proved the feasibility of the N-doping co-activation process.
The influence of temperature on BB was measured by a TGA test under N2 protection, as shown in Figure 3A. The TGA curves corresponding to three main stages of weight loss ranged from room temperature to 1200 °C. The first stage occurred at room temperature to 275 °C and was caused by evaporation and loss of residual water from physical surfaces and internal pores [34,35]. The second stage, which started at 275 °C and ended at 500 °C, was the most significant stage of weight loss. It can be explained by the fact that the main oxygenated component of BB was lignocellulose, which can be cracked into gases and tar at higher temperatures. With the removal of these pyrolytic substances, weight loss in the second stage was induced. Therefore, 500 °C was selected as the carbonization temperature for CBB preparation, and the yield was 20.28% at this time. As for the third stage (500–1200 °C), the TG curve was relatively stable, indicating that there was no obvious weight loss phenomenon. However, when the temperature exceeded 1000 °C (especially when it was close to 1200 °C), CBB exhibited a slight mass change, which may have been caused by the pyrolysis of some minerals [36,37].
The FT-IR spectra of the functional groups of the samples were analyzed, as shown in Figure 2B. The broad band at around 3450 and 2930 cm−1 represented the stretching vibrations of hydroxyl functional groups (O-H) and -CH, -CH2, and -CH3 groups [38]. The band around 1730 cm−1 represented the stretching vibration of C=O. The peak at 1630 cm−1 represented the axial deformation of the carbonyl group. The band at around 1375 and 1420 cm−1 represented the C-H symmetric bending vibration of the methyl group and the deformation vibration of methylene [38,39]. In common with many biomass materials [34,35,38,39], the bands at around 1047–1159 cm−1 represented the tensile vibrations of C-O from alcohols, phenols, acids, or esters. After N-doping co-activation, a new vibration peak appeared at 1690 cm−1, which indicated that the C=N was formed on N-BBB [20].
The crystal structure of the samples was tested by XRD, as shown in Figure 3C. The peaks at 17° and 23° represent the cellulose from the lignocellulose of raw BB [32,33]. The irregular peaks represent minerals that can be interpreted as inorganic salts. These peaks became sharper and more distinct after carbonization. After direct activation and N-doped co-activation, these peaks were significantly weakened, which can be explained by the removal of a large quantity of soluble mineral salts during the washing process. In addition, we noted that the diffraction peaks of CBB, BBB, and N-BBB were in the range of 10–30° and 38–45°, indicating that the prepared biochars had the local structure of typical carbon materials; that is, they contained both amorphous carbon and a 2D graphite planar structure [40].
The presence of defects in the carbon was determined by Raman spectra, as shown in Figure 3D. Two typical peaks obtained from the results include the D-band with amorphous carbon at around 1339 ± 7 cm−1 and the G-band with graphitic carbon at around 1583 ± 3 cm−1 [38,39,40]. To measure the degree of defect and disorder in the carbons, the intensity ratio of the D-band and G-band (ID/IG) was used as an important index. The ID/IG value of CBB was 1.02. After activation, the ID/IG value of BBB and N-BBB were 1.26 and 1.20, which indicated that more amorphous carbon structures were generated in the biochars.
The specific surface area and porosity of the samples were tested by N2 adsorption-desorption isotherms, as shown in Figure 4, and the data are shown in Table 1. The specific surface area and the total pore volume of CBB were 49.5 m2/g and 0.0222 cm3/g, respectively, which were not enough to support CBB as a porous biochar for adsorption. Therefore, further activation treatment is needed to greatly improve its properties and enhance its application performance. After direct activation and N-doping co-activation, the specific surface areas of BBB and N-BBB were 2502.3 and 2292.7 m2/g, respectively, and their total pore volumes were 1.1389 and 1.0356 cm3/g, respectively. Compared with CBB, the values were greatly improved, which indicated the success and effectiveness of the further activation treatment. Moreover, both BBB and N-BBB showed typical type IV isotherms with a slight H3 hysteresis loop, indicating that some mesoporous structures existed in the prepared biochars [39,40]. The volumes of micropores in BBB and N-BBB were 1.1118 and 1.0170 cm3/g, respectively, accounting for 97.6% and 98.2% of the total pore volume, respectively, indicating that they were porous biochars with mainly micropore structure and partly mesoporous structure [24,25]. Pore size distribution was used to further analyze the porosity of the samples. The results based on the NLDFT method also showed that both BBB and N-BBB had microporous and mesoporous structures. The BJH and H-K methods were used to study the pore distribution, which not only re-analyzed the multi-pore structure of BBB and N-BBB but also further proved the existence of both mesoporous and microporous structures.
In addition, it is noteworthy that the specific surface area and total pore volume of N-BBB were smaller than those of BBB. There are three possible reasons for this. First, the addition of the nitrogen source urea in the activation process will compete with the activator NaOH for the contact area with the carbon precursor, thus affecting the specific surface area and pore structure of N-BBB. Second, it may be caused by the chemical reaction between the ammonia gas generated by the nitrogen source urea at high temperature and carbon dioxide produced in the activation process and water vapor, thus consuming part of the activator and reducing the activation efficiency. Third, the nitrogen source urea and activator simultaneously etched the surface of the carbon material and generated more gases [29,30,31], which diffused into the carbon precursor, resulting in partial fragmentation, affecting the pore structure and resulting in the reduction in the specific surface area and total pore volume.
The surface chemical and electronic states of the samples were determined by XPS spectra, as shown in Figure 5. Both BBB and N-BBB contained mainly C, O, and N elements, which was consistent with the EDS. The high-resolution C1s spectra of BBB and N-BBB showed three classical peaks at 283.87–283.88, 284.82–284.98, and 287.74–288.22 eV corresponding to C-C, C-O, and C=O, respectively [38,39,40]. The high-resolution O1s spectra of BBB and N-BBB both had three peaks at 530.74–530.93, 532.36–532.43, and 533.76–534.28 eV corresponding to C=O, C-O, and -OH, respectively [38,39,40]. The high-resolution N1s spectra of ITGB and MITGB both had peaks at 397.41–37.42 and 399.33–399.38 eV, corresponding to pyridinic-N and pyrrolic-N, respectively [38,39,40]. In addition, N-BBB had a unique peak corresponding to graphite-N at 403.53 eV [20], indicating that the N-doped co-activation method had indeed successfully doped N into N-BBB.

2.3. Results of Adsorption Performances

2.3.1. Adsorption Kinetics

Adsorption kinetics describes the adsorption capacity of the adsorbent at different initial solution concentrations as a function of contact time [32,33,34,35]. Therefore, the effect of time on the adsorption of MO and Cr6+ by BBB and N-BBB at a temperature of 303 K was explored, as shown in Figure 6. Whether the adsorbent was BBB or N-BBB, or the adsorbate was MO or Cr6+, all the adsorption process trends were similar. The adsorption capacities increased sharply in the first 30 min, and then gradually reached equilibrium at 60 min. The extension of contact time did not further significantly improve the adsorption capacities. It can be speculated that the adsorption capacities would increase with the increase in the initial concentration of the solution, and the high concentration of solution promoted the adsorption process to some certain extent [40]. In order to study the control mechanism of reactions in the process of adsorption, three common adsorption kinetic models, Lagergren’s PFK model based on surface physical adsorption [40], Ho–McKay’s PSK model based on chemical adsorption [39], and Weber–Morris’s IPD model based on molecular diffusion [38] were used to analyze the experimental data, as shown in Table 2.
The PFK correlation coefficients R2 of BBB for MO ranged from 0.9780 to 0.9944; while the Qe.cat values (611.8, 720.4, and 800.3 mg/g) were lower than the Qe (627.9, 737.2, and 836.9 mg/g) obtained from the experiments. The PFK for Cr6+ ranged from 0.9550 to 0.9741; while the Qe.cat values were 85.3, 109.2, and 130.4 mg/g for different initial concentrations, and were again lower than Qe (93.8, 119.6, and 141.1 mg/g). To summarize, the PFK model was not the best kinetic to describe the whole adsorption process. The IPD correlation coefficients R2 of BBB for MO ranged from 0.5822 to 0.7517, indicating that the adsorption process of MO by BBB may not be affected by particle diffusion. For Cr6+, on the contrary, the correlation coefficients R2 ranged from 0.9138 to 0.9280, which indicated the adsorption process had a particle diffusion behavior. Ho–McKay’s PSK model was used to fit the data; the PSK correlation coefficients R2 of BBB ranged from 0.9848 to 0.9997 for MO and from 0.9913 to 0.9919 for Cr6+. Meanwhile, the Qe.cat values (664.5, 769.0, and 844.2 mg/g for MO, and 98.4, 122.3, and 143.6 mg/g for Cr6+) agreed with those obtained from experiments, which showed the applicability of PSK in the adsorption process.
When the models were used to fit with the data of N-BBB, the PFK correlation coefficients R2 were 0.9721–0.9952 for MO and 0.9318–0.9506 for Cr6+. The Qe.cat values were 631.3, 732.9, and 817.1 mg/g for MO, and 89.7, 122.4 and 153.0 mg/g for Cr6+, which were both lower than the Qe values (644.4, 755.1 and 858.3 mg/g for MO, and 99.5, 135.0, and 169.1 mg/g for Cr6+) obtained from experiments. It can be speculated that the PFK model may have played a role in the adsorption process, although the role was not dominant. For the IPD model, the correlation coefficients R2 were 0.5471–0.7830 for MO and 0.9432–0.9746 for Cr6+, also indicating that the adsorption process of N-BBB for Cr6+ had a particle diffusion behavior and was largely unaffected by particle diffusion for MO. The PSK coefficients R2 were 0.9873–0.9989 for MO and 0.9765–0.9869 for Cr6+, indicating the PSK model was more suitable to describe the adsorption process. Moreover, it was found that, with the increase in the initial concentration of the solution, the rate constants k2 of BBB were also increased for both MO and Cr6+, indicating that the adsorption rate gradually accelerated, and the adsorption rate was faster at a higher concentration. However, the rate constants k2 of N-BBB were different. For MO, N-BBB showed a similar trend to BBB. At the same time, the rate constants k2 of N-BBB for Cr6+ decreased gradually, which indicated that, with the increase in initial concentration, the adsorption rate gradually slowed, and the adsorption rate was slower at a higher concentration.
According to the results, we inferred that the adsorption processes of BBB and N-BBB for MO and Cr6+ may be mainly chemical reactions (while physical adsorption and particle diffusion also had certain effects on the adsorption process), and the adsorption behaviors between adsorbent and adsorbate through transfer, exchange, or sharing to form chemisorption bonds, may control the adsorption rate [40,41].

2.3.2. Adsorption Isotherms

The effect of concentration on the adsorption capacity of the adsorbent is usually determined by investigating the adsorption isotherms. The effects of different initial concentrations of solution on the adsorption process of BBB and N-BBB were studied at a temperature of 303 K, and the results are shown in Figure 7. Increasing the initial concentration of MO or Cr6+ solution was beneficial to the forward process of adsorption.
Subsequently, the Langmuir and Freundlich isotherm models were used to analyze the experimental data, and the results are shown in Table 3. The Langmuir isotherm model is often used to describe the adsorption process of homogeneous molecules [42], while the Freundlich isotherm model is often used to study the heterogeneous multilayer adsorption process [43]. When the adsorbate was MO, the Langmuir isotherm correlation coefficients R2 were 0.9327 for BBB and 0.9041 for N-BBB, which indicated that the adsorption processes were not uniform single-layer. The Qm of BBB and N-BBB for MO were 860.3 and 887.4 mg/g, higher than Qe, indicating that the prepared biochars had higher adsorption capacities for MO. Moreover, the KL of N-BBB for MO was slightly bigger than that of BBB, which showed that N-BBB had a faster MO adsorption rate. The Freundlich isotherm correlation coefficients R2 for MO were 0.9967 for BBB and 0.9883 for N-BBB. The nF values of BBB and N-BBB for MO were 7.75 and 7.51, bigger than 1.0, indicating that this adsorption model of fitting MO was appropriate. As for Cr6+, the Langmuir isotherm correlation coefficients R2 were 0.9949 for BBB and 0.9822 for N-BBB, which also showed that the uniform single-layer adsorption was not suitable to describe the adsorption process. When the Freundlich isotherm was used to fit the data, the correlation coefficients R2 were 0.9976 for BBB and 0.9959 for N-BBB; while the nF values of BBB and N-BBB were 3.76 and 2.77. The results indicated that the adsorption processes were non-uniform multilayer [42,43,44].

2.3.3. Adsorption Thermodynamics

The influence of temperature (293, 303, and 313 K) on adsorption by BBB and N-BBB is shown in Figure 8. On increasing the temperature from 293 K to 313 K, the adsorption capacities of BBB for MO and Cr6+ increased from 712.5 to 764.8 mg/g and from 105.8 to 124.1 mg/g, respectively, while the adsorption capacities of N-BBB for MO and Cr6+ increased from 726.7 to 773.7 mg/g and from 112.6 to 148.3 mg/g, respectively. Obviously, raising the temperature increased the adsorption capacities of MO and Cr6+—that is, a high-temperature environment promoted the adsorption processes by biochars (BBB and N-BBB).
The experimental data were analyzed by thermodynamics formulas, and the parameters are shown in Table 4. All ΔG values were negative, indicating that the adsorption occurred spontaneously, both with BBB (−5.86, −6.19, and −6.55 for MO and −0.27, −0.61, and −0.73 for Cr6+) and N-BBB (−5.93, −6.29, and −6.60 for MO and −0.43, −0.93, and −1.23 for Cr6+) [39,40]. The thermodynamic enthalpy values of ΔH of adsorption of MO were 4.29 and 3.82 kJ/mol for BBB and N-BBB, respectively, and the ΔH values of adsorption of Cr6+ were 6.47 and 11.23 kJ/mol for BBB and N-BBB, respectively, which further confirmed the endothermic property of the adsorption process [38]. In addition, the positive values of thermodynamic ΔS (34.61 and 33.27 Jmol−1 K−1 for MO, 23.01 and 39.78 J mol−1 K−1 for Cr6+) indicated that the randomness and chaos of the interface between the porous biochars and solutions increased with the increase in temperature [40].

2.3.4. Effect of pH

In general, pH affects the adsorption process by changing the charge properties of the adsorbent and the adsorbate [32,33,34,35]. MO has two chemical structures, basic and acidic, and whether the chromophore was anthraquinone or azo bond depends on the pH of the solution [45]. The adsorption of MO by biochars was investigated in the pH range of 2 to 10, and the results are shown in Figure 9A. With the increase in pH, the adsorption capacity of both BBB and N-BBB decreased, which can be explained by the electrostatic attraction between the surface charge of the biochars and the ionic charge of the anionic dye MO. Compared with BBB, N-BBB had higher electronegativity due to the N-doping. When the pH value was less than 6, the affinity of the dye increased, and the negative charge on the surface of the biochar could be used as the active site to generate a strong electrostatic attraction with the dye in solution. Conversely, when the pH was higher than 6 (particularly 8), the protonation of the dye was gradually weakened, and the biochars with negative active sites on the surface were not conducive to the adsorption of anionic dyes due to electrostatic repulsion, which reduced the adsorption amount of MO. Cr6+ also possessed various forms at different pH levels [4,5,20], where it exists as HCrO4 and Cr2O72− when the pH is lower than 6.5, and as Cr2O72− and CrO42− when the pH is higher than 6.5. With the increase in pH, the adsorption capacities of both BBB and N-BBB decreased, and the maximum adsorption capacities were at a pH value of 2. This can be explained by the fact that the biochars (BBB and N-BBB) had a positive surface charge at a pH less than pHpzc (4.46 for BBB and 4.52 for N-BBB). When the pH was higher than 6, the adsorption capacities decreased sharply, which may be due to electrostatic repulsion. When the pH was in the range of 8 to 10, the adsorption became relatively stable, which indicated that electrostatic interaction was not the only process affecting adsorption performance [20].

2.4. Results of Cycle Tests

The recyclability of adsorbents is an important parameter for evaluating the practical performance of biochars [38,39,40]; thus, the five-cycle performance of the biochars was investigated, and the results are shown in Figure 10. The removal rates of MO and Cr6+ by BBB and N-BBB decreased with an increase in cycle number. This can be explained as follows: on the one hand, with the treatment of the cycling experiment, adsorbed organic pollutants formed by-products on the surface of biochars [38,39]; while on the other hand, with the increase in the carbonization regeneration process, the structure of the biochars became more fragile, which further affected its regeneration. At the same time, after five cycles, the removal rate of MO and Cr6+ by BBB and N-BBB remained above 80%, indicating that they had good stability and regeneration ability.

2.5. Probable Mechanism Analysis

In this work, the N-BBB exhibited good removal ability of MO and Cr6+ in water solutions affected by many factors (Figure 11). Firstly, the large specific surface area and high total pore volume (2292.7 m2/g and 1.0356 cm3/g) of N-BBB provided many adsorption sites for the adsorption pollutant; therefore, it can be speculated that pore filling may play an important role in the adsorption process. The results based on the kinetics and isotherm showed that the adsorption process was heterogeneous multilayer adsorption with a chemical reaction, which indicated that the chemical binding force will also promote the adsorption process. In addition, from the test results with FT-IR and XPS, it can be speculated that many unsaturated functional groups containing carbon and oxygen on the biochar surface will produce hydrogen-bond interaction with the pollutant model. Moreover, FT-IR and Raman test results also showed that the prepared biochars contained aromatic rings and sp2 hybridized carbon with graphite structure, and the π bond in these structures may also have π–π interaction with aromatic rings in pollutants to enhance the adsorption capacity. The charged pollutants in the appropriate pH environment formed a strong electrostatic attraction with N-BBB, which further promoted the adsorption process. Experiments under different conditions, such as the initial concentration of the solution and the temperature of the adsorption process, show that these also affect the adsorption process. To summarize, in addition to the experimental conditions, the pore filling, π–π interaction, H-bond interaction, and electrostatic attraction supported the excellent performance of N-BBB.

2.6. Comparison

Adsorption capacity is an important parameter for evaluating the practical performance of an adsorbent; thus, the adsorption capacities of BBB and N-BBB are compared with other biochars, as shown in Table 5. The adsorption capacities of CBB for MO and Cr6+ were only 25.2 and 10.4 mg/g. After activation and N-doping co-activation, the adsorption capacities of BBB and N-BBB to MO and Cr6+ were significantly enhanced. In addition, compared with other biochars, the adsorption capacities of BBB and N-BBB were not low, which fully indicated that the prepared biochars had great potential and application prospects in the treatment of water pollutants.

3. Materials and Methods

3.1. Materials and Reagents

Birch bark (BB), Betula Mandshurica Nakai, obtained from the campus of Jilin Agricultural University (Changchun, China) in 2022, was washed with deionized water, dried at 80 °C for 12 h, and crushed. Urea, NaOH, H2SO4, HCl, and ethanol were purchased from Beijing Chemical Works (Beijing, China) and used without further purification.
Diphenylcarbazide (CAS: 140-22-7), Methyl orange (MO, CAS: 547-58-0), and Potassium dichromate (Cr6+, CAS: 7778-50-9) were supplied by Aladdin Chemical (Shanghai) Co., Ltd. (Shanghai, China) and the structural formulas are shown in Figure S1 (Supplementary Material).

3.2. Preparation of Biochars

BB was carbonized at 500 °C for 60 min with a heating rate of 10 °C/min under the protection of a nitrogen atmosphere to obtain CBB. BB was used by mixing with NaOH and Urea at a ratio of 1:4:1; meanwhile, 1.0 g BB was sufficiently ground with 4.0 g NaOH. The two mixtures were both heated at 700 °C for 60 min. After cooling to room temperature, the activated mixtures were washed with HCl and deionized water until reaching a natural pH value and dried at 180 °C for 12 h. At last, the samples including CBB, BBB, and N-BBB were kept in a desiccator prior to subsequent experiments.

3.3. Adsorption Performances

In a batch adsorption experiment, 0.05 g/L BBB or N-BBB was added to a flask containing pollutant solutions (MO or Cr6+). The flask was placed in a constant temperature shaker at 150.0 RPM in the dark. After the adsorption process reached equilibrium, the suspension was centrifuged, and the supernatant was diluted with deionized water. The concentration of the solution was determined with an Agilent Cary-300 UV-vis spectrophotometer. The adsorption capacities of samples were calculated by Equation (1):
Q e = ( C 0 C e ) × V m
where Qe (mg/g) represents the adsorption capacity of the sample, Ce represents the equilibrium concentrations of the solution, C0 (mg/L) represents the initial concentrations of the solution, m (g) represents the mass of the samples, and V (L) represents the volume of the solutions.
The pollutant solutions were prepared at different concentrations (50, 100, and 200 mg/L). A total of 0.05 g/L BBB or N-BBB was dispersed into flasks containing MO or Cr6+ solutions and shaken at 150 RPM in the dark at 303 K. Then, the concentrations of the solutions were determined at preset time intervals. The pseudo-first-order kinetic (PFK, Equation (2)), the pseudo-second-order kinetic (PSK, Equation (3)), and the intra-particle diffusion model (IPD, Equation (4)) were used to analyze the adsorption kinetic data, shown as follows:
ln ( Q e Q t ) = ln Q e k 1 t
t Q t = 1 k 2 Q t 2 + t Q e
Q t = k 3 t 0.5 + C
where Qt represents the adsorption capacity of the sample at different time points t, C represents the thickness of the boundary layer, k1 represents the PFK adsorption kinetic rate constant, k2 represents the PSK adsorption kinetic rate constant, and k3 denotes the IPD adsorption kinetic rate constant.
The pollutant solutions at different initial concentrations (50, 100, 150, 200, and 250 mg/L) were prepared and used to test the adsorption isotherm at 303 K. After adsorption saturation, the absorbances of the solutions were measured using a UV-Vis spectrophotometer. The adsorption isotherm data were investigated using the Langmuir isotherm model (Equation (5)) and Freundlich isotherm model (Equation (6)), as follows:
C e Q e = C e Q m + 1 Q m K L
ln Q m = 1 n ln C e + ln K F
where Qm (mg/g) represents the maximum adsorption capacity of the sample calculated by the adsorption isotherm model, KL represents the Langmuir adsorption isotherm constant, and KF represents the Freundlich adsorption isotherm constant.
The effect of temperature (293, 303, and 313 K) on the adsorption capacity of the samples was investigated at an initial concentration of 100 mg/L BBB or 0.05 g/L N-BBB. The thermodynamic parameters were analyzed to describe the effect of temperature on the adsorption process. The calculation equations were as follows:
ln K T = Δ H R T + Δ S R
K T = Q e C e
Δ G = Δ H T Δ S
where ΔS represents the thermodynamic parameters’ standard entropy, ΔG represents the standard free Gibbs energy, ΔH represents the standard enthalpy, and R represents the gas constant (8.314 J/K·mol).
The variation in the adsorption capacity of the samples with the pH (2, 4, 6, 8, and 10) was also investigated. The solutions were adjusted to different pH values by HCl and NaOH.

3.4. Cycle Tests

In each cycle, 1.0 g/L BBB or N-BBB was placed into a flask containing the organic pollutant solutions at a concentration of 100 mg/L at 303 K. After the adsorption of pollutants (Figure S2, Supplementary Material), the samples, BBB/MO, BBB/Cr6+, N-BBB/MO, and N-BBB/Cr6+, were collected and washed with deionized water. Then, the recycled samples were carbonized for 60 min at 600 °C under the protection of a nitrogen atmosphere. The re-carbonized samples were re-used as fresh adsorbent in the next cycle.

4. Conclusions

In this work, BB was used for the first time to prepare porous biochars via N-doping co-activation. The specific surface area and total pore volume of N-BBB were 2292.7 m2/g and 1.0356 cm3/g, respectively, which proved the feasibility of N-doping co-activation in pore-forming. The large specific surface area and the high total pore volume played a substantial role in the adsorption process. EDS, FT-IR, and XPS were used to characterize the samples, and the results indicated that nitrogen doping was successfully completed. In an experiment using the synthetic dye MO and heavy metal Cr6+ as the pollutant models, N-BBB showed good removal ability. The adsorption capacity of N-BBB remained above 80% after five regenerations, which fully proved the stability of regeneration. Moreover, the excellent adsorption performance of N-BBB may have been influenced by pore filling, π–π interaction, H-bond interaction, and electrostatic attraction. This study not only provided a biochar adsorbent with excellent performance but also verified the feasibility of the one-step nitrogen-doping co-activation method to prepare high-performance biochars. In the future, we will continue to study this N-doped co-activation method and explore its application to other types of biomass to further develop more biochars with better performance.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms232314618/s1.

Author Contributions

Conceptualization, S.C. and Y.S. (Yingjie Su); methodology, Y.S. (Yuqing Shi); software, M.J.; validation, Y.S. (Yingjie Su); formal analysis, Y.S. (Yuqing Shi); investigation, M.J.; resources, Y.S. (Yuqing Shi); data curation, M.J.; writing—original draft preparation, S.C.; writing—review and editing, S.C.; visualization, Y.S. (Yingjie Su); supervision, S.C.; project administration, Y.S. (Yingjie Su); funding acquisition, Y.S. (Yingjie Su). All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by [Jilin Scientific and Technological Development Program] grant number [20220203108SF] And The APC was funded by [20220203108SF].

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

We acknowledge the support of the Jilin Scientific and Technological Development Program (20220203108SF). This research was also supported by the China National College Students Innovation and Entrepreneurship Project.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhou, L.; Wu, Y.; Zhang, S.; Li, Y.; Gao, Y.; Zhang, W.; Tian, L.; Li, T.; Du, Q.; Sun, S. Recent development in microbial electrochemical technologies: Biofilm formation, regulation, and application in water pollution prevention and control. J. Water Process Eng. 2022, 49, 103135. [Google Scholar] [CrossRef]
  2. Yan, C.; Qu, Z.; Wang, J.; Cao, L.; Han, Q. Microalgal bioremediation of heavy metal pollution in water: Recent advances, challenges, and prospects. Chemosphere 2022, 286, 131870. [Google Scholar] [CrossRef] [PubMed]
  3. Zamora-Ledezma, C.; Negrete-Bolagay, D.; Figueroa, F.; Zamora-Ledezma, E.; Ni, M.; Alexis, F.; Guerrero, V.H. Heavy metal water pollution: A fresh look about hazards, novel and conventional remediation methods. Environ. Technol. Innov. 2021, 22, 101504. [Google Scholar] [CrossRef]
  4. Xu, H.; Fan, Y.; Xia, X.; Liu, Z.; Yang, S. Effect of Ginkgo biloba leaves on the removal efficiency of Cr(VI) in soil and its underlying mechanism. Environ. Res. 2023, 216, 114431. [Google Scholar] [CrossRef]
  5. Njoya, O.; Zhao, S.; Shen, J.; Kong, X.; Gong, Y.; Wang, B.; Kang, J.; Chen, Z. Acetate improves catalytic performance for rapid removal of Cr(VI) by sodium borohydride in aqueous environments. Sep. Purif. Technol. 2022, 301, 122051. [Google Scholar] [CrossRef]
  6. Li, X.Q.; Zhang, Q.H.; Ma, K.; Li, H.M.; Guo, Z. Identification and determination of 34 water-soluble synthetic dyes infoodstuff by high performance liquid chromatography–diode arraydetection–ion trap time–of–flight tandem mass spectrometry. Food Chem. 2015, 182, 316–326. [Google Scholar] [CrossRef]
  7. Moghadas, M.R.S.; Motamedi, E.; Nasiri, J.; Naghavi, M.R.; Sabokdast, M. Proficient dye removal from water using biogenic silver nanoparticles prepared through solid-state synthetic route. Heliyon 2020, 6, e04730. [Google Scholar] [CrossRef]
  8. Mandal, B.; Ray, S.K. Synthesis of interpenetrating network hydrogel from poly(acrylic acid-co-hydroxyethyl methacrylate) and sodium alginate: Modeling and kinetics study for removal of synthetic dyes from water. Carbohyd. Polym. 2013, 98, 257–269. [Google Scholar] [CrossRef]
  9. Baig, U.; Kashif Uddin, M.; Gondal, M.A. Removal of hazardous azo dye from water using synthetic nano adsorbent: Facile synthesis, characterization, adsorption, regeneration and design of experiments. Colloids Surf. A 2020, 584, 124031. [Google Scholar] [CrossRef]
  10. Nidheesh, P.V.; Zhou, M.; Oturan, M.A. An overview on the removal of synthetic dyes from water by electrochemical advanced oxidation processes. Chemosphere 2018, 197, 210–227. [Google Scholar] [CrossRef]
  11. Ambika; Kumar, V.; Jamwal, A.; Kumar, V.; Singh, D. Green bioprocess for degradation of synthetic dyes mixture using consortium of laccase-producing bacteria from Himalayan niches. J. Environ. Manag. 2022, 310, 114764. [Google Scholar] [CrossRef] [PubMed]
  12. Zaafouri, Z.; Batot, G.; Nieto-Draghi, C.; Coasne, B.; Bauer, D. Impact of adsorption kinetics on pollutant dispersion in water flowing in nanopores: A Lattice Boltzmann approach to stationary and transient conditions. Adv. Water Resour. 2022, 162, 104143. [Google Scholar] [CrossRef]
  13. Tee, G.T.; Gok, X.Y.; Yong, W.F. Adsorption of pollutants in wastewater via biosorbents, nanoparticles and magnetic biosorbents: A review. Environ. Res. 2022, 212, 113248. [Google Scholar] [CrossRef]
  14. Zhou, W.; Zhang, W.; Cai, Y. Enzyme-enhanced adsorption of laccase immobilized graphene oxide for micro-pollutant removal. Sep. Purif. Technol. 2022, 294, 121178. [Google Scholar] [CrossRef]
  15. Ma, Y.; Li, Y.; Zhao, X.; Zhang, L.; Wang, B.; Nie, A.; Mu, C.; Xiang, J.; Zhai, K.; Xue, T.; et al. Lightweight and multifunctional super-hydrophobic aramid nanofiber/multiwalled carbon nanotubes/Fe3O4 aerogel for microwave absorption, thermal insulation and pollutants adsorption. J. Alloys Compd. 2022, 919, 165792. [Google Scholar] [CrossRef]
  16. Sellaoui, L.; Gomez-Aviles, A.; Dhaouadi, F.; Bedia, J.; Bonilla-Petriciolet, A.; Rtimi, S.; Belver, C. Adsorption of emerging pollutants on lignin-based activated carbon: Analysis of adsorption mechanism via characterization, kinetics and equilibrium studies. Chem. Eng. J. 2023, 452, 139399. [Google Scholar] [CrossRef]
  17. Qiu, B.; Shao, Q.; Shi, J.; Yang, C.; Chu, H. Application of biochar for the adsorption of organic pollutants from wastewater: Modification strategies, mechanisms and challenges. Sep. Purif. Technol. 2022, 300, 121925. [Google Scholar] [CrossRef]
  18. Zhang, B.; Wang, M.; Qu, J.; Zhang, Y.; Liu, H. Characterization and mechanism analysis of tylosin biodegradation and simultaneous ammonia nitrogen removal with strain Klebsiella pneumoniae TN-1. Bioresour. Technol. 2021, 336, 125342. [Google Scholar] [CrossRef]
  19. Wu, J.; Wang, T.; Liu, Y.; Tang, W.; Geng, S.; Chen, J. Norfloxacin adsorption and subsequent degradation on ball-milling tailored N-doped biochar. Chemosphere 2022, 303, 135264. [Google Scholar] [CrossRef]
  20. Qu, J.; Zhang, X.; Liu, S.; Li, X.; Wang, S.; Feng, Z.; Wu, Z.; Wang, L.; Jiang, Z.; Zhang, Y. One-step preparation of Fe/N co-doped porous biochar for chromium(VI) and bisphenol a decontamination in water: Insights to co-activation and adsorption mechanisms. Bioresour. Technol. 2022, 361, 127718. [Google Scholar] [CrossRef]
  21. Li, J.; Lin, Q.; Luo, H.; Fu, H.; Wu, L.; Chen, Y.; Ma, Y. The effect of nanoscale zero-valent iron-loaded N-doped biochar on the generation of free radicals and nonradicals by peroxydisulfate activation. J. Water Process Eng. 2022, 47, 102681. [Google Scholar] [CrossRef]
  22. Tan, X.; Zhu, S.; Wang, R.; Chen, Y.; Showf, P.; Zhang, F.; Ho, S. Role of biochar surface characteristics in the adsorption of aromatic compounds: Pore structure and functional groups. Chin. Chem. Lett. 2021, 32, 2939–2946. [Google Scholar] [CrossRef]
  23. Zhang, J.; Huang, D.; Shao, J.; Zhang, X.; Yang, H.; Zhang, S.; Chen, H. Activation-free synthesis of nitrogen-doped biochar for enhanced adsorption of CO2. J. Clean. Prod. 2022, 355, 131642. [Google Scholar] [CrossRef]
  24. Liew, R.; Azwar, E.; Yek, P.; Lim, X.; Cheng, C.; Ng, J.; Jusoh, A.; Lam, W.; Ibrahim, M.; Ma, N.; et al. Microwave pyrolysis with KOH/NaOH mixture activation: A new approach to produce micro-mesoporous activated carbon for textile dye adsorption. Bioresour. Technol. 2018, 266, 1–10. [Google Scholar] [CrossRef]
  25. Chen, S.; Zhang, B.; Xia, Y.; Chen, H.; Chen, G.; Tang, S. Influence of mixed alkali on the preparation of edible fungus substrate porous carbon material and its application for the removal of dye. Colloids Surf. A 2021, 609, 125675. [Google Scholar] [CrossRef]
  26. Wei, M.; Marrakchi, F.; Yuan, C.; Cheng, X.; Jiang, D.; Zafar, F.F.; Fu, Y.; Wang, S. Adsorption modeling, thermodynamics, and DFT simulation of tetracycline onto mesoporous and high-surface-area NaOH-activated macroalgae carbon. J. Hazard. Mater. 2022, 425, 127887. [Google Scholar] [CrossRef] [PubMed]
  27. Alfatah, T.; Mistar, E.M.; Supardan, M.D. Porous structure and adsorptive properties of activated carbon derived from Bambusa vulgaris striata by two-stage KOH/NaOH mixture activation for Hg2+ removal. J. Water Process Eng. 2021, 43, 102294. [Google Scholar] [CrossRef]
  28. Kamran, U.; Park, S. Tuning ratios of KOH and NaOH on acetic acid-mediated chitosan-based porous carbons for improving their textural features and CO2 uptakes. J. CO2 Util. 2020, 40, 101212. [Google Scholar] [CrossRef]
  29. Yang, X.; Wang, Q.; Lai, J.; Cai, Z.; Lv, J.; Chen, X.; Chen, Y.; Zheng, X.; Huang, B.; Lin, G. Nitrogen-doped activated carbons via melamine-assisted NaOH/KOH/urea aqueous system for high performance supercapacitors. Mater. Chem. Phys. 2020, 250, 123201. [Google Scholar] [CrossRef]
  30. Hou, Z.; Tao, Y.; Bai, T.; Liang, Y.; Huang, S.; Cai, J. Efficient Rhodamine B removal by N-doped hierarchical carbons obtained from KOH activation and urea-oxidation of glucose hydrochar. J. Environ. Chem. Eng. 2021, 9, 105757. [Google Scholar] [CrossRef]
  31. Wang, K.; Xu, M.; Gu, Y.; Gu, Z.; Fan, Q.H. Symmetric supercapacitors using urea-modified lignin derived Ndoped porous carbon as electrode materials in liquid and solid electrolytes. J. Power Sources 2016, 332, 180–186. [Google Scholar] [CrossRef] [Green Version]
  32. Jin, Y.; Zhang, B.; Chen, G.; Chen, H.; Tang, S. Combining biological and chemical methods to disassemble of cellulose from corn straw for the preparation of porous carbons with enhanced adsorption performance. Int. J. Biol. Macromol. 2022, 209, 315–329. [Google Scholar] [CrossRef] [PubMed]
  33. Chen, S.; Xia, Y.; Zhang, B.; Chen, H.; Chen, G.; Tang, S. Disassembly of lignocellulose into cellulose, hemicellulose, and lignin for preparation of porous carbon materials with enhanced performances. J. Hazard. Mater. 2021, 408, 124956. [Google Scholar] [CrossRef] [PubMed]
  34. Zhang, B.; Jin, Y.; Qi, J.; Chen, H.; Chen, G.; Tang, S. Porous carbon materials based on Physalis alkekengi L. husk and its application for removal of malachite green. Environ. Technol. Innov. 2021, 21, 101343. [Google Scholar] [CrossRef]
  35. Xia, Y.; Jin, Y.; Qi, J.; Chen, H.; Chen, G.; Tang, S. Preparation of biomass carbon material based on Fomes fomentarius via alkali activation and its application for the removal of brilliant green in wastewater. Environ. Technol. Innov. 2021, 23, 101659. [Google Scholar] [CrossRef]
  36. Fan, C.; Yan, J.; Huang, Y.; Han, X.; Jiang, X. XRD and TG-FTIR study of the effect of mineral matrix on the pyrolysis and combustion of organic matter in shale char. Fuel 2015, 139, 502–510. [Google Scholar] [CrossRef]
  37. Yan, J.; Jiang, X.; Han, X.; Liu, J. A TG–FTIR investigation to the catalytic effect of mineral matrix in oil shaleon the pyrolysis and combustion of kerogen. Fuel 2013, 104, 307–317. [Google Scholar] [CrossRef]
  38. Zhang, B.; Jin, Y.; Huang, X.; Tang, S.; Chen, H.; Su, Y.; Yu, X.; Chen, S.; Chen, G. Biological self-assembled hyphae/starch porous carbon composites for removal of organic pollutants from water. Chem. Eng. J. 2022, 450, 138264. [Google Scholar] [CrossRef]
  39. Xia, Y.; Zhang, B.; Guo, Z.; Tang, S.; Su, Y.; Yu, X.; Chen, S.; Chen, G. Fungal mycelium modified hierarchical porous carbon with enhanced performance and its application for removal of organic pollutants. J. Environ. Chem. Eng. 2022, 10, 108699. [Google Scholar] [CrossRef]
  40. Chen, X.; Yu, G.; Chen, Y.; Tang, S.; Su, Y. Cow dung-based biochar materials prepared via mixed base and its application in the removal of organic pollutants. Int. J. Mol. Sci. 2022, 23, 10094. [Google Scholar] [CrossRef]
  41. Chauhdary, Y.; Hanif, M.A.; Rashid, U.; Bhatti, I.A.; Anwar, H.; Jamil, Y.; Alharthi, F.A.; Kazerooni, E.A. Effective removal of reactive and direct dyes from colored wastewater using low-cost novel bentonite nanocomposites. Water 2022, 14, 3604. [Google Scholar] [CrossRef]
  42. Lin, L.; Li, L.; Xiao, L.; Zhang, C.; Li, X.; Pervez, M.N.; Zhang, Y.; Nuruzzaman, M.; Mondal, M.I.H.; Cai, Y.; et al. Adsorption behaviour of reactive blue 194 on raw Ramie Yarn in palm oil and water media. Materials 2022, 15, 7818. [Google Scholar] [CrossRef] [PubMed]
  43. El-Sayed, N.S.; Salama, A.; Guarino, V. Coupling of 3-Aminopropyl sulfonic acid to cellulose nanofifibers for effificient removal of cationic dyes. Materials 2022, 15, 6964. [Google Scholar] [CrossRef] [PubMed]
  44. Shirendev, N.; Bat-Amgalan, M.; Kano, N.; Kim, H.-J.; Gunchin, B.; Ganbat, B.; Yunden, G. A natural zeolite developed with 3-Aminopropyltriethoxylane and adsorption of Cu(II) from aqueous media. Appl. Sci. 2022, 12, 11344. [Google Scholar] [CrossRef]
  45. Jiao, Y.; Xu, L.; Sun, H.; Deng, Y.; Zhang, T.; Liu, G. Synthesis of benzxazine-based nitrogen-doped mesoporous carbon spheres for methyl orange dye adsorption. J. Porous Mater. 2017, 24, 1565–1574. [Google Scholar] [CrossRef]
  46. Zhang, B.; Wu, Y.; Cha, L. Removal of methyl orange dye using activated biochar derived from pomelo peel wastes: Performance, isotherm, and kinetic studies. J. Disper. Sci. Technol. 2018, 41, 1561298. [Google Scholar] [CrossRef]
  47. Zhang, H.; Li, R.; Zhang, Z. A versatile EDTA and chitosan bi-functionalized magnetic bamboo biochar for simultaneous removal of methyl orange and heavy metals from complex wastewater. Environ. Pollut. 2022, 293, 118517. [Google Scholar] [CrossRef]
  48. Ouedrhiri, A.; Himi, M.A.; Youbi, B.; Lghazi, Y.; Bahar, J.; Haimer, C.E.; Aynaou, A.; Bimaghra, I. Biochar material derived from natural waste with superior dye adsorption performance. Mater. Today. Proc. 2022, 66, 259–267. [Google Scholar] [CrossRef]
  49. Hou, Y.; Liang, Y.; Hu, H.; Tao, Y.; Zhou, J.; Cai, J. Facile preparation of multi-porous biochar from lotus biomass for methyl orange removal: Kinetics, isotherms, and regeneration studies. Bioresour. Technol. 2021, 329, 124877. [Google Scholar] [CrossRef]
  50. Aichour, A.; Zaghouane-Boudiaf, H.; Khodja, H.D. Highly removal of anionic dye from aqueous medium using a promising biochar derived from date palm petioles: Characterization, adsorption properties and reuse studies. Arab. J. Chem. 2022, 15, 103542. [Google Scholar] [CrossRef]
  51. Li, Y.; Chen, X.; Liu, L.; Liu, P.; Zhou, Z.; Huhetaoli; Wu, Y.; Lei, T. Characteristics and adsorption of Cr(VI) of biochar pyrolyzed from landfill leachate sludge. J. Anal. Appl. Pyrolysis 2022, 162, 105449. [Google Scholar] [CrossRef]
  52. Zhen, Z.; Duan, X.; Tie, J. One-pot synthesis of a magnetic Zn/iron-based sludge/biochar composite for aqueous Cr(VI) adsorption. Environ. Technol. Innov. 2022, 28, 102661. [Google Scholar] [CrossRef]
  53. Xu, D.; Sun, T.; Jia, H.; Sun, Y.; Zhu, X. The performance and mechanism of Cr(VI) adsorption by biochar derived from Potamogeton crispus at different pyrolysis temperatures. J. Anal. Appl. Pyrolysis 2022, 167, 105662. [Google Scholar] [CrossRef]
  54. Yi, Y.; Wang, X.; Zhang, Y.; Ma, J.; Ning, P. Adsorption properties and mechanism of Cr(VI) by Fe2(SO4)3 modified biochar derived from Egeria najas. Colloids Surf. A 2022, 645, 128938. [Google Scholar] [CrossRef]
  55. Kuang, Q.; Liu, K.; Wang, Q.; Chang, Q. Three-dimensional hierarchical pore biochar prepared from soybean protein and its excellent Cr(VI) adsorption. Sep. Purif. Technol. 2023, 304, 122295. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of the preparation.
Figure 1. Schematic diagram of the preparation.
Ijms 23 14618 g001
Figure 2. SEM images of (A) BB, (B) CBB, (C) BBB, and (E) N-BBB. EDS mapping of (D) BBB and (F) N-BBB.
Figure 2. SEM images of (A) BB, (B) CBB, (C) BBB, and (E) N-BBB. EDS mapping of (D) BBB and (F) N-BBB.
Ijms 23 14618 g002
Figure 3. (A) TGA curves of BB. (B) FT-IR spectra and (C) XRD of BB, CBB, BBB, and N-BBB. (D) Raman spectra of CBB, BBB, and N-BBB.
Figure 3. (A) TGA curves of BB. (B) FT-IR spectra and (C) XRD of BB, CBB, BBB, and N-BBB. (D) Raman spectra of CBB, BBB, and N-BBB.
Ijms 23 14618 g003
Figure 4. (A) N2 adsorption-desorption isotherms of samples. Pore distribution of samples based on (B) NLDFT method, (C) BJH method, and (D) H-K method.
Figure 4. (A) N2 adsorption-desorption isotherms of samples. Pore distribution of samples based on (B) NLDFT method, (C) BJH method, and (D) H-K method.
Ijms 23 14618 g004
Figure 5. XPS spectra of (A) BBB and (B) N-BBB. The C1s, O1s, and N1s high-resolution spectra of BBB (CE) and N-BBB (FH).
Figure 5. XPS spectra of (A) BBB and (B) N-BBB. The C1s, O1s, and N1s high-resolution spectra of BBB (CE) and N-BBB (FH).
Ijms 23 14618 g005
Figure 6. PFK, PSK, and IPD plots of MO for (A) BBB and (B) N-BBB, and plots of Cr6+ for (C) BBB and (D) N-BBB at 303 K.
Figure 6. PFK, PSK, and IPD plots of MO for (A) BBB and (B) N-BBB, and plots of Cr6+ for (C) BBB and (D) N-BBB at 303 K.
Ijms 23 14618 g006
Figure 7. Adsorption isotherms of MO and Cr6+ for BBB (A,B) and N-BBB (C,D) at 303 K.
Figure 7. Adsorption isotherms of MO and Cr6+ for BBB (A,B) and N-BBB (C,D) at 303 K.
Ijms 23 14618 g007
Figure 8. Adsorption thermodynamics of BBB and N-BBB for (A) MO and (B) Cr6+.
Figure 8. Adsorption thermodynamics of BBB and N-BBB for (A) MO and (B) Cr6+.
Ijms 23 14618 g008
Figure 9. Effect of pH on the adsorption capacities of MO and Cr6+ onto (A) BBB and (B) N-BBB (inset: Zeta potential).
Figure 9. Effect of pH on the adsorption capacities of MO and Cr6+ onto (A) BBB and (B) N-BBB (inset: Zeta potential).
Ijms 23 14618 g009
Figure 10. Cycling stability tests of BBB and N-BBB for (A) MO and (B) Cr6+.
Figure 10. Cycling stability tests of BBB and N-BBB for (A) MO and (B) Cr6+.
Ijms 23 14618 g010
Figure 11. Probable mechanisms analysis for N-BBB removal of MO and Cr6+.
Figure 11. Probable mechanisms analysis for N-BBB removal of MO and Cr6+.
Ijms 23 14618 g011
Table 1. The data of N2 adsorption-desorption for CBB, BBB, and N-BBB.
Table 1. The data of N2 adsorption-desorption for CBB, BBB, and N-BBB.
SamplesSBET (m2/g)Vmicro (cm3/g)Vtotal (cm3/g)
CBB49.50.01870.0222
BBB2502.31.11181.1389
N-BBB2292.71.01701.0356
SBET, Vmicro, and Vtotal represent the BET specific surface area, the volume of micropores, and the total pore volume.
Table 2. Fitting parameters of adsorption kinetic models for MO and Cr6+ at 303 K.
Table 2. Fitting parameters of adsorption kinetic models for MO and Cr6+ at 303 K.
AdsorbatesAdsorbentsModelsParametersC0 (mg L−1)
50100200
MOBBB Qe (mg/g)627.9737.2836.9
PFKk1 (min−1)0.00280.00740.0058
Qe.cat (mg/g)611.8720.4800.3
R20.98800.99440.9780
PSKk2 (g mg−1 min−1)0.00030.00030.0005
Qe.cat (mg/g)664.5769.0844.2
R20.98480.98960.9997
IPDk3 (mg g−1 min−0.5)27.1223.6219.90
C350.1497.9628.2
R20.64640.58220.7517
N-BBB Qe (mg/g)644.4755.1858.3
PFKk1 (min−1)0.00440.00540.0051
Qe.cat (mg/g)631.3732.9817.1
R20.99520.98850.9721
PSKk2 (g mg−1 min−1)0.00040.00040.0005
Qe.cat (mg/g)672.1777.9862.8
R20.98730.99800.9989
IPDk3 (mg g−1 min−0.5)19.422.020.9
C448.7530.3637.3
R20.54710.65490.7830
Cr6+BBB Qe (mg/g)93.8119.6141.1
PFKk1 (min−1)0.00140.00170.0020
Qe.cat (mg/g)85.3109.2130.4
R20.97410.96140.9550
PSKk2 (g mg−1 min−1)0.00070.00080.0008
Qe.cat (mg/g)98.4122.3143.6
R20.99230.99190.9913
IPDk3 (mg g−1 min−0.5)6.26.97.3
C22.741.459.1
R20.91380.92800.9279
N-BBB Qe (mg/g)99.5135.0169.1
PFKk1 (min−1)0.00150.00170.0017
Qe.cat (mg/g)89.7122.4153.0
R20.95000.95060.9318
PSKk2 (g mg−1 min−1)0.00070.00070.0006
Qe.cat (mg/g)102.2135.9170.6
R20.98430.98690.9765
IPDk3 (mg g−1 min−0.5)6.27.39.6
C27.750.558.4
R20.95370.94320.9746
Table 3. Fitting parameters of adsorption isotherm models for MO and Cr6+ at 303 K.
Table 3. Fitting parameters of adsorption isotherm models for MO and Cr6+ at 303 K.
AdsorbatesAdsorbentsTypesParameters
MOBBBLangmuirQm (mg/g)860.3
KL (L/mg)0.1345
R20.9327
FreundlichKF (mg g−1(L mg−1)1/n)431.5
nF7.75
R20.9967
N-BBBLangmuirQm (mg/g)887.4
KL (L/mg)0.1364
R20.9041
FreundlichKF (mg g−1(L mg−1)1/n)437.1
nF7.51
R20.9883
Cr6+BBBLangmuirQm (mg/g)169.8
KL (L/mg)0.0264
R20.9949
FreundlichKF (mg g−1(L mg−1)1/n)34.9
nF3.76
R20.9976
N-BBBLangmuirQm (mg/g)226.2
KL (L/mg)0.0163
R20.9822
FreundlichKF (mg g−1(L mg−1)1/n)25.7
nF2.77
R20.9959
Table 4. Fitting adsorption thermodynamic parameters for MO and Cr6+.
Table 4. Fitting adsorption thermodynamic parameters for MO and Cr6+.
AdsorbentsAdsorbatesT (K)G
(kJ/mol)
H
(kJ/mol)
S
(J mol−1 K−1)
BBBMO293−5.864.2934.61
303−6.19
313−6.55
Cr6+293−0.276.4723.01
303−0.61
313−0.73
N-BBBMO293−5.933.8233.27
303−6.29
313−6.60
Cr6+293−0.4311.2339.78
303−0.93
313−123
Table 5. Comparison of BBB and N-BBB to MO and Cr6+ with other biochars.
Table 5. Comparison of BBB and N-BBB to MO and Cr6+ with other biochars.
AdsorbentsQe for MO (mg/g)Qe for Cr6+ (mg/g)References
BBB836.9141.1This work
N-BBB858.3169.1This work
Pomelo peel biochar147.9-[46]
Magnetic bamboo biochar305.4-[47]
Date seeds biochar334.0-[48]
Lotus root biochar449.0-[49]
Date palm petioles biochar461.0-[50]
Landfill leachate sludge biochar-17.5[51]
Zn/iron-based sludge/biochar-27.0[52]
Potamogeton crispus biochar-34.4[53]
Egeria najas biochar-138.8[54]
Soybean protein biochar-489.7[55]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Su, Y.; Shi, Y.; Jiang, M.; Chen, S. One-Step Synthesis of Nitrogen-Doped Porous Biochar Based on N-Doping Co-Activation Method and Its Application in Water Pollutants Control. Int. J. Mol. Sci. 2022, 23, 14618. https://doi.org/10.3390/ijms232314618

AMA Style

Su Y, Shi Y, Jiang M, Chen S. One-Step Synthesis of Nitrogen-Doped Porous Biochar Based on N-Doping Co-Activation Method and Its Application in Water Pollutants Control. International Journal of Molecular Sciences. 2022; 23(23):14618. https://doi.org/10.3390/ijms232314618

Chicago/Turabian Style

Su, Yingjie, Yuqing Shi, Meiyi Jiang, and Siji Chen. 2022. "One-Step Synthesis of Nitrogen-Doped Porous Biochar Based on N-Doping Co-Activation Method and Its Application in Water Pollutants Control" International Journal of Molecular Sciences 23, no. 23: 14618. https://doi.org/10.3390/ijms232314618

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