Next Article in Journal
Spatial and Temporal Differences and Influencing Factors of Eco-Efficiency of Cultivated Land Use in Main Grain-Producing Areas of China
Previous Article in Journal
Barriers of and Possibilities for Recycling of Single-Use Take-Away Food and Beverage Packaging: Evidence from Lithuanian Market
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Iron- and Nitrogen-Modified Biochar for Nitrate Adsorption from Aqueous Solution

by
Sohrab Haghighi Mood
1,*,
Manuel Raul Pelaez-Samaniego
2,
Yinglei Han
1,3,
Kalidas Mainali
1,4 and
Manuel Garcia-Perez
1
1
Department of Biological Systems Engineering, Washington State University, Pullman, WA 99164, USA
2
Department of Applied Chemistry and Production Systems, Faculty of Chemical Sciences, University of Cuenca, Cuenca 010107, Ecuador
3
Department of Bioresource and Environmental Security, Sandia National Laboratories, 7011 East Avenue, Livermore, CA 94550, USA
4
US Department of Agriculture, Agricultural Research Service, Eastern Regional Research Center, Sustainable Biofuels and Co-Products Research Unit, Wyndmoor, PA 19038, USA
*
Author to whom correspondence should be addressed.
Sustainability 2024, 16(13), 5733; https://doi.org/10.3390/su16135733
Submission received: 16 May 2024 / Revised: 21 June 2024 / Accepted: 1 July 2024 / Published: 4 July 2024

Abstract

:
Nutrient pollution poses a significant global environmental threat, and addressing this issue remains an ongoing challenge. Biochar has been identified as a potential adsorbent for environmental remediation. However, raw biochar has a low nitrate adsorption capacity; thus, biochar modification is necessary for targeted environmental applications. This work explored and compared the performance of Fe-doped, N-doped, and N-Fe-co-doped biochars from Douglas fir toward nitrate removal from an aqueous solution. A central composite experimental design was used to optimize processing variables, maximizing the surface area and nitrate adsorption capacity. Proximate analysis, elemental composition, gas physisorption, XPS, SEM, TEM, FTIR, and XRD were used to characterize the biochar’s properties. Pyrolysis under NH3 gas generated more pores in biochar than conventional pyrolysis. Doping biochar with N and Fe improved nitrate adsorption capacity from aqueous solutions. The maximum nitrate adsorption capacity of Fe-N-doped biochar produced at 800 °C was 20.67 mg g−1 in sorption tests at pH 3.0. The formation of N-containing functional groups and Fe oxides on the biochar surface enhanced the nitrate removal efficiency of N-Fe biochar. The results indicate that biochar’s adsorption capacity for NO3 is largely affected by the solution’s pH and biochar’s surface chemistry. Electrostatic attraction is the primary mechanism for nitrate adsorption.

1. Introduction

Nutrient pollution has become a critical environmental issue of global concern. Although nitrogen (N) is an essential element in aquatic ecosystems, excessive use of nitrogenous fertilizers and disposal of untreated industrial wastes can lead to elevated levels of N in natural water systems, thus impairing water quality [1,2,3,4]. High concentrations of N and P in water lead to eutrophication, threatening the ecosystem and human health [5]. Nitrate (NO3) is a monovalent anion with lone electrons and strong electronegativity [6] that is highly soluble in water, making it one of the most widespread groundwater contaminants [7]. High NO3 concentrations in water can cause some types of cancer and a fatal disorder called methemoglobinemia [8,9]. The U.S. Environmental Protection Agency (US EPA) set the maximum contamination level of NO3 in drinking water to be 10 mg L−1 [10], which is approximately equivalent to the World Health Organization (WHO) guideline (i.e., 50 mg L−1 as NO3, equivalent to ~11 mg L−1 as nitrate nitrogen NO3-N) [11]. However, the concentrations of NO3 in groundwater in some regions can surpass 80 mg L−1 [12] and 200 mg L−1 of NO3 [13], by far exceeding the WHO’s guidelines. Therefore, there is an urgent necessity to develop effective and low-cost solutions to remove NO3 from water sources [14]. Groundwater contamination by NO3 can originate from various sources, such as wastewater treatment plants, poorly managed septic systems, industrial activities, leaky urban sewers, urban runoff, fertilizers, and animal waste [15]. For example, a study conducted in California showed that, for over 50 years, NO3 from fertilizer and animal waste had contaminated the Tulare Lake Basin and Salinas Valley aquifers [16]. Similarly, Adimalla et al. conducted a study on the chemistry, distribution, and potential health risks of NO3 in groundwater in a semi-urban area of South India [12]. The study findings indicated that 43.3% of groundwater samples exceeded the safety limit of 45 mg L−1 according to Indian guidelines [12]. The authors suggested that the main sources of NO3 in groundwater are non-point sources, such as fertilizers, animal waste, and poor sanitation facilities [12].
Different technologies and processes have been employed to remove NO3 from water, including reverse osmosis, electrodialysis, ion exchange, biological denitrification, chemical denitrification, and adsorption [17]. Biological denitrification is a cost-effective and environmentally friendly method [18,19], but the process could be long, limiting its use [20]. Additionally, this method is sensitive to environmental conditions [21]. Ion exchange is also effective, but it requires careful waste brine disposal and costly ion exchange resins. Reverse osmosis, although effective, is also an expensive method [17]. Chemical denitrification is not feasible at a large scale due to the excessive operational costs involved [22]. Electrodialysis has some drawbacks that still need to be considered. Depending on the level and type of pollutants present in the feed water, a pre-treatment stage may be required. Additionally, pH regulation may be necessary for the water obtained after the electrodialysis process [23]. Adsorption offers the potential to be an efficient and cheaper method for removing NO3. Its easy operation makes adsorption one of the best technologies for removing pollutants from water [24,25]. In the case of agricultural runoff, adsorption seems to be the only viable path. Operational parameters such as solution pH, contact time, temperature, adsorbent dosage, and NO3 concentration should be carefully controlled for efficient NO3 adsorption [6]. Different materials, such as metal oxides/hydroxides, carbon-based materials, organic polymer adsorbents, and agricultural wastes, are used as NO3 adsorbents [6].
Biochar is one of the main products of biomass pyrolysis under oxygen-limited atmospheres. Biochar’s unique properties, such as a high surface area, porous structure, and tunable surface chemistry, make it a suitable adsorbent for a wide range of contaminants. [4,25]. Positively charged surfaces containing exchangeable anions, multivalent metal ions, and/or hydrogen atoms can be active sites for NO3 adsorption. However, despite the biochar’s tunable properties, most commercially available biochars have a deficient adsorption capacity. The lack of suitable surface functional groups and the required porosity results in poor NO3 adsorption. Therefore, biochar modification is essential for NO3 adsorption to obtain tuned structural and chemical properties. The surface charge of biochar can be largely influenced by the pretreatment of biomass, the pyrolysis conditions, and the post-treatment of biochar. Therefore, these factors need to be considered and controlled when producing biochar for various applications [26]. Biochar produced via conventional pyrolysis under inert gas has high oxygen functional groups [17,27] and typically has a negative surface charge at neutral pH, which is not favorable for the adsorption of anions (e.g., NO3) [28,29]. Different modification strategies have been used to improve biochar’s anion adsorption capacity. Positively charged primary functional groups such as amides, aromatic amines, and pyridinic groups could be suitable adsorption sites for negatively charged nitrate ions [30,31]. Introducing heteroatoms to biochar carbon polyaromatic ring systems can improve electronic properties, including readily available lone pairs of electrons, durability, thermal stability, charging and polarizability, and conductivity [32]. Heteroatom-containing functional groups can alter the surface properties of biochar since these groups are generally polar and possess sites for hydrogen bonding, ion–dipole, and dipole–dipole interactions [33]. In contrast, hydrophobic carbon-rich surfaces of biochar are suitable sites for non-polar moieties [34].
N doping can be an effective way to introduce these basic functional groups on biochar surfaces [31,35,36]. The role of N in biochar to create/modify functional groups relies on the fact that N has more electronegativity than C, which enables electron transfer from neighboring C, resulting in a high charge density of C atoms [36,37], as shown in previous studies. Yoo et al. produced N-doped activated carbon to remove NO3 from an aqueous solution and found that N-doping increased the adsorption capacity of activated carbon from 0.38 to 0.75 mmol g−1 [38]. One method for introducing N functional groups involves modifying the environment employed for producing biochar. Pyrolysis of biomass under NH3 gas was shown to introduce basic N-containing functional groups on the biochar surface due to the reduction in the electron-withdrawing acidic groups and the increase in π electron-dense and oxygen-containing basic groups [39]. This method produced N-doped chars from anaerobically digested fiber with an improved affinity towards anionic pollutants [28]. The increase in the adsorption capacity of the modified biochar was explained by the formation of basic functional groups [31,35,40]. It has also been indicated that introducing metal oxides on biochar surfaces might lead to a positively charged surface that can increase the NO3 adsorption rate [41]. Several methods have been used to introduce Fe to biochar structures [42,43,44]. Fe can form different Fe oxides on biochar surfaces that improve NO3 adsorption [42,43,44].
The previously mentioned studies indicate that N doping and Fe doping have been used individually to remove NO3 from water. However, the combined impact of biochar N and Fe doping has not been addressed in the literature, to the best of the authors’ knowledge. Therefore, the aim of this work is to evaluate the influence of N doping, Fe doping, and Fe-N co-doping on the adsorption capacity of biochar obtained from Douglas fir wood. To the best of the authors’ knowledge, it is the first time that the combined effect of N and Fe on NO3 adsorption has been investigated. Fe-N-co-doped biochars were produced through one-step NH3 pyrolysis of FeCl3-loaded Douglas fir. Response Surface Methodology (RSM) with central composite design was used to examine the effects of biochar processing parameters and Fe content on the adsorption capacity towards NO3 in an aqueous solution.

2. Materials and Methods

2.1. Biochar Production

Douglas fir wood (DF) residues were used as feedstock for biochar production. DF was air-dried to reduce the moisture content below 8 wt.%, ground, and sieved to obtain particles that passed through a 0.6 mm sieve. The following four types of biochars were produced using N2 or NH3 gases: (1) biochar produced through pyrolysis under N2 gas (herein referred to as DF biochar), (2) biochar produced through pyrolysis under NH3 gas (i.e., N-DF biochar), (3) biomass impregnated with FeCl3 solution pyrolyzed under N2 gas (i.e., Fe-DF biochar); and (4) biomass impregnated with FeCl3 pyrolyzed under NH3 gas (i.e., Fe-N-DF biochar). All biochars were produced in duplicate. Figure 1 depicts the processes to produce the biochars. Regarding the fourth type of biochar (Fe-N-DF), biochar was first produced at one temperature (800 °C), metal loading (1 g of FeCl3 per 15 g of DF), and pyrolysis time (1 h) to compare the NO3 adsorption performance with other types of biochars. Then, the process (for type 4 biochar) was optimized to maximize the biochar surface area and NO3 adsorption capacity.
The DF biochar (the first type of biochar) was produced through pyrolysis of the DF particles under N2 gas at 800 °C in a quartz tube furnace reactor that was 1000 mm long with a 50 mm external diameter (Thermo Scientific Lindberg/Blue Hinged tube furnace HTF55000 Series, Asheville, NC, USA) with three heating zones (controlled independently) that work at similar temperatures. The materials were placed inside the tube using three quartz combustion boats. Approximately 5 g of ground DF was placed in each boat (i.e., ~15 g in total). The heating-up period was performed under N2. Briefly, the DF particles were in contact with N2 for 30 min at 25 °C. Afterward, the temperature was increased from 25 °C to 800 °C at a heating rate of 10 °C min−1 and kept at 800 °C for one hour. A flow rate of 500 mL min−1 of N2 was employed. Then, the biochar produced was cooled down to 100 °C or below under N2 before removal. No recovery of the pyrolysis liquid by-product was conducted.
The second type of biochar, N-DF biochar, was produced following the same procedure, but instead of N2, NH3 gas was used (at a flow rate of 600 mL min−1). For the third type of biochar, Fe-doped biochar (Fe-DF biochar), 15 g of DF, 1 g of FeCl3 (Sigma Aldrich St. Louis, MO, USA), and 200 mL of water were mixed using a magnetic stirrer at 60 °C for 6 h. The mixture was then filtered using filter paper to separate Fe-containing DF from FeCl3 solution (after filtration, some FeCl3 content remains in the solution), and then Fe-containing DF was dried at 80 °C for 48 h. The remaining solid was subjected to pyrolysis under N2 gas. Finally, the Fe- and N-co-doped biochar (Fe-N-DF biochar) was prepared through pyrolysis of Fe-containing DF under NH3 gas in a similar way to that employed for the N-DF biochar. In all cases, the final biochar products were washed thoroughly with ionized water until a neutral pH was achieved.

2.2. Char Characterization

Proximate analysis: Proximate analysis was performed in triplicates, using a thermogravimetric analyzer (TGA), SDTA851e (Mettler Toledo, Columbus, OH, USA), to determine char moisture, fixed carbon, volatiles, and ash content. Biochar was heated under N2 gas in a crucible from 25 to 120 °C and held at 120 °C for 3 min to determine moisture content. The biochar was then heated to 950 °C under N2 gas (50 mL min−1) and held for 5 min to assess the volatile matter content. Then, the biochar was cooled down to 450 °C. Ash was measured as the remaining mass after biochar burning at 600 °C under an oxygen gas environment for 8 min [28].
Elemental analysis: Elemental analysis was carried out using a TRUSPEC-CHN® (LECO, St. Joe, MI, USA) elemental analyzer to measure total nitrogen (N), Hydrogen (H), and carbon (C), following the process described previously [28]. Oxygen (O) was computed by difference. These analyses were conducted in triplicates.
Gas physisorption analysis: N2 and CO2 adsorption isotherms were measured at 76.85 K and 273 K on a TriStar II PLUS Surface Area and Porosity Analyzer (Micromeritics, Norcross, GA, USA). The micropore volumes were analyzed from the CO2 adsorption using the Dubinine–Radushkevich (DR) equation. A non-localized density functional theory (NLDFT) was employed using commercial software (MicroActive™ v.1.01, Micromeritics, Norcross, GA, USA) to determine the pore size distribution. Density functional theory also provided an independent assessment of the volume of pores [45,46].
X-ray photoelectron spectroscopy (XPS): XPS was carried out using a Kratos AXIS ULTRADLD XPS (Manchester, UK) system equipped with an Al Ka X-ray source and a 165 mm mean radius electron energy hemispherical analyzer. Vacuum pressure was kept below 3 × 10−9 torr during the acquisition.
Char morphology: Scanning electron microscope (SEM) imaging analysis was performed using a Tescan Vega3 instrument (Brno, Czech Republic) combined with energy dispersive spectroscopy (EDS). Biochar particles were mounted on a stub and gold-coated before visualization. SEM and EDS were applied to examine the structure and surface characteristics of the char before and after adsorption. A transmission electron microscope (TEM) was used to study biochar’s surface morphology and micro and nanostructure. For the test, each biochar sample was ground into powder. A suspension of distilled (DI) water and fine powder samples were prepared and dispersed onto copper grids. Imaging was conducted at 200 kV under vacuum conditions with an FEI Tecnai G2 20 Twin [47].
X-ray powder diffraction (XRD): The crystallinity of biochars was studied using XRD (Miniflex 600 benchtop X-ray diffractometer Rigaku, Tokyo, Japan) with Cu K α radiation and operated at 40 kV, 15 mA, with 0.01 degree-steps and a scanning rate of 0.5° min−1. The scan range of interest for this analysis was 20–100°.
Fourier Transform Infrared Spectroscopy (FTIR): FTIR analysis was conducted to identify the functional groups on the biochars’ surfaces. FTIR spectra were obtained using a Shimadzu IRPrestige 21 spectrometer (Tokyo, Japan) equipped with a MIRacle single reflection ATR Ge probe with the spectra recording from 900 to 3900 cm−1.

2.3. Nitrate Adsorption Study

Adsorption of NO3 was carried out to obtain isotherms to evaluate the adsorption efficiency of the four types of biochars. These experiments were conducted in duplicate. In each case, 0.1 g of biochar was mixed with 35 mL of nitrate solution with different concentrations of nitrate–nitrogen (NO3-N) ranging from 10 to 30 mg L−1 in 50 mL tubes at room temperature (25 °C). Nitrate solutions were prepared by dissolving NaNO3 (Fisher Scientific,Fair Lawn, NJ, USA) in deionized water. Then, the tubes were shaken at 130 rpm in a mechanical shaker (orbital shaker, Lab-Line Instruments, Inc., Melrose Park, IL, USA) for 24 h. The samples were then filtered using 0.45 μm filters to measure the corresponding nitrate equilibrium concentration. The pH of the solutions was measured using a pH meter (Mettler Toledo, SevenEasy S20) before and after NO3 adsorption. NO3 adsorption capacities were determined by measuring the initial and the final concentrations of the NO3 in the aqueous solution, which were evaluated by the Chromotropic acid method [48]. To explore the effect of pH on NO3 adsorption, the pH of the solution was set to three different values (i.e., 3, 7, and 8). The pH of the solutions was adjusted using 0.1 M HCl or 0.1 M NaOH.

2.4. Experimental Design

Multiple variables influence the pyrolysis of Fe-impregnated Douglas fir. Therefore, the process needs to be optimized to identify independent variables (pyrolysis temperatures, pyrolysis time, and metal loading) and maximize biochar surface area and NO3 adsorption capacity. A central composite design (CCD) was applied to optimize the preparation conditions of Fe-N-DF biochar. Statistical software Design Expert, version 13 (Stat-Ease, Inc., Minneapolis, MN, USA), was used in our analysis. The independent factors and their coded levels for the CCD are shown in Table 1.

3. Results and Discussion

3.1. Biochar Production and Characterization

Fe and N significantly affected the pyrolysis process, biochar yield, and biochar structure. The yields of biochars produced (at 800 °C for 1 h in all cases) were 21.1, 20.5, 28.2, and 24.3 wt.% for DF, N-DF, Fe-DF, and Fe-N-DF biochar, respectively. The presence of FeCl3 contributed to an increase in biochar yields, which is consistent with previous findings [28,29,49]. The results of the proximate and elemental analysis are presented in Table 2 and Table 3, respectively. Impregnation of DF with FeCl3 solution resulted in biochar with higher ash content and lower fixed carbon than DF biochar. As shown in Table 3, N content in Fe-N-DF biochar is higher (8 wt.%) than in N-DF biochar (6.6 wt.%), which results from the presence of Fe because Fe in biochar catalyzes the formation of the N dopant [28,50]. The N and O contents increased after Fe-N co-doping, indicating that Fe and N were introduced into the biochar matrix and modified the chemical composition of biochar. According to Luo et al., introducing N-containing functional groups can occur through the reaction of NH3 with oxygen-containing functional groups [51]. Elemental analysis results suggest that introducing Fe to DF increased the relative O content in the resulting biochar, possibly due to the formation of O-containing functional groups, such as Fe oxides [52,53].
Biochar’s surface areas and pore volumes obtained with CO2 and N2 adsorption are shown in Table 4. The N2 adsorption method is performed at a temperature of −196 °C. This usually leads to lower surface area values for materials with micropores (<2 nm) because of the kinetic diameter of the gas molecule (0.36 nm) [25,54]. Slow diffusion of N2 at −196 °C can limit the use of N2 in the range between 0.7 and 1.47 nm. A combination of CO2 and N2 adsorption should be used to characterize biochar [25,55]. N-DF biochar exhibited the highest specific surface area and pore volume. The results suggest that pyrolysis under NH3 gas generates more pores in biochar structure than conventional pyrolysis under inert gas (e.g., N2). Previous studies have reported that N doping through NH3 treatment can increase the specific surface area of biochar [51]. Pyrolysis under ammonia gas results in carbon skeleton etching by gas–solid reactions, producing biochar with a high surface area [50,56]. At around 500 °C, NH3 decomposes into ˙NH2, ˙NH, H atoms, and N atoms. The radicals produced in this process react with active substances such as C=O or –OH to create N-containing functional groups. These radicals also take part in gas–solid reactions and etch the surface of the biochar, which increases the pore volume and specific surface area through the pathway shown in equation (1) [57].
(Cbiomass) + NH3 → C* + H* + NH* + NH*2 → H2 + C-NH + C-NH2
The surface areas of Fe-DF and Fe-N-DF biochars obtained from CO2 and N2 adsorption are lower than that of pristine biochar, which can be attributed to the blockage of pores by Fe oxides and the resulting decrease in the surface area [58]. Pores can be occupied by Fe nanoparticles and reduce the surface area of biochar [59]. The pore size distribution of biochar is shown in Figure 2, indicating that the pore structure of biochar (N-DF) has been improved by pyrolysis under NH3 gas. Moreover, most of the pores in all types of biochar have a width between 0.30 and 0.70 nm. The CO2 and N2 adsorption isotherms obtained from gas physisorption analyses are shown in Figure 3 and Figure 4. The isotherm in Figure 3 shows mainly the presence of micropores. IUPAC conventions offer a standardized method for categorizing pore sizes and gas sorption isotherms, providing valuable insights into the correlation between porosity and sorption. The adsorption–desorption isotherm obtained from the N2 method provides information about porosity, specific surface area, and pore size distribution [60]. Figure 3c for the Fe-DF biochar clearly shows a typical IUPAC type IV isotherm with an H3 hysteresis loop, indicating that the Fe-DF biochar is mainly mesoporous [61,62]. The isotherms presented in Figure 3a,b (e.g., N-DF) can be categorized as type I isotherms, which suggests that biochar is considered a microporous solid [63]. Moreover, the isotherm shown in Figure 3d can be classified as a type II isotherm. CO2 adsorption isotherms (Figure 4) indicate that the amount of CO2 adsorbed at p/p0 = 0.03 also increased when pyrolysis of DF was conducted under an NH3 environment, showing an increase in micropores in the biochar.
The combined effect of pyrolysis temperature, metal loading, and pyrolysis time on the surface area of Fe-N-co-doped Douglas fir biochar is shown in Figure 5. We tested experimental optimization strategies with a central composite experimental design for biochar in which we varied the metal loading by 2 to 6 g per 15 g of Douglas fir wood, pyrolysis time (between 30 and 90 min), and the final pyrolysis temperature (between 600 and 800 °C) as input variables to identify the conditions resulting in a product with the highest surface area. As seen in Figure 5, the highest surface area can be achieved at the pyrolysis temperature of 698 °C (~700 °C), metal loading of 2 g per 15 g of DF, and the pyrolysis time of 90 min. In Figure 5, it can be observed that the pyrolysis time has a minimal effect on the biochar surface area.
FTIR was used to study the characteristics of functional groups in biochars (Figure 6). Bands between 3723 and 3176 cm−1 were attributed to –OH and N–H functional groups. However, these peaks cannot be observed in Figure 6 since the –OH group disappears in biochars produced at high temperatures (e.g., 800 °C). The rise observed at 1080 cm−1 can be attributed to C–O/C–N stretching vibration. The peak at around 1220 cm−1 could be attributed to C–N stretching. Peaks between 1400 and 1680 cm−1 could be ascribed to C=O, C=C aromatic, and C=N stretching. The N-DF and Fe-N-DF biochars were produced under NH3 gas; therefore, there is a high possibility of C=N formation. Moreover, peaks between 2800 and 2900 cm−1 can be observed, which might be attributed to aliphatic C–H [64,65].
The SEM images (Figure 7a,b) show that fine particles were dispersed on the surface of Fe-N-DF biochar (produced at 800 °C), which could be due to the formation of metal oxides during the pyrolysis. The EDS result shows the presence of primarily N, Fe, and O, in addition to C (the main constituent) in biochar structures (see Figure 7c). The presence of N and Fe suggests that these elements were successfully doped in biochar structures. TEM images of Fe-N-DF (Figure 7d,e) showed irregular and amorphous morphology on the surface. Small spherical particles can be seen in the biochar structure, which could be due to the formation of nanoparticles of Fe oxides, showing that Fe was embedded in the carbon structure. TEM images also show the presence of nanoparticles between 20 and 50 nm. This result suggests that in addition to the Fe integrated into the biochar structure, some Fe also precipitated on the surface. Iron was detected on the biochars’ surface (Figure 7c,f).

3.2. Nitrate Adsorption

Nitrate adsorption studies were conducted at three different pH levels (3, 7, and 8), using four types of produced biochars. Four different isotherms, including Langmuir, Freundlich, Langmuir–Freundlich, and Redlich–Peterson (Equations (1)–(4)), were applied to explain the adsorption characteristics observed for NO3 of each material. In the Langmuir equation, K (L mg−1) and Q (mg g−1) denote the Langmuir bonding term, which is associated with interaction energies and Langmuir maximum capacity, respectively, and Ce (mg L−1) is the equilibrium solution concentration of the sorbate. K (mg(1−n) Ln·g−1) and n (dimensionless) in Freundlich represent the Redlich affinity coefficient and the Freundlich linearity constant, respectively. K (Ln mg−n) and n denote the Langmuir–Freundlich affinity parameter and heterogeneity index, respectively. K (L g−1), a (Ln mg−n), and n in the Redlich–Peterson equation represent Redlich–Peterson isotherm constants. The Langmuir model corresponds to monolayer adsorption on a homogeneous adsorbent surface with no interactions between the adsorbed molecules [66,67]. The Freundlich model considers a heterogeneous sorption process with a non-uniform distribution of adsorption heat, which can be applied to multilayer adsorption [68]. The Langmuir–Freundlich model is a combined isotherm model of Langmuir and Freundlich isotherms. Thus, the Langmuir–Freundlich model can be used to model both homogeneous and heterogeneous binding surfaces [69]. Likewise, the Redlich–Peterson model can be applied to either homogeneous or heterogeneous systems since this model is a hybrid isotherm that features Langmuir and Freundlich isotherms [70,71]. Equations (2)–(5) describe the mentioned models.
q = K   Q   C e 1 + K   C e Langmuir
q = K   C e n Freundlich
q = K   Q   C e n 1 + K   C e n Langmuir–Freundlich
q = K   C e 1 + a   C e n Redlich–Peterson
The adsorption isotherms of DF, N-DF, Fe-DF, and Fe-N-DF biochars (produced at 800 °C, metal loading of 1 g of FeCl3 to 15 g of DF, and pyrolysis time of 1 h), at pH 7, are shown in Figure 8. The maximum NO3 adsorption capacity for Fe-N-DF biochar at pH 7 was 9.37 mg g−1, which is approximately two-fold higher than that of pristine biochar (4.82 mg g−1). Generally, the surface of pristine biochars is negatively charged, limiting its binding affinity toward NO3. The results suggest that introducing N and Fe to the biochar structure improves NO3 adsorption on the biochar surface. It also appears that although N-DF biochar exhibited the highest surface area, the role of N and Fe functional groups in Fe-N-DF biochar is more important than the surface area for NO3 adsorption. The isotherms in Figure 8 indicate that the Langmuir–Freundlich and Redlich–Peterson models provided the best fit. The results also show that decreasing pH caused improved NO3 adsorption capacity, which might be attributed to the protonation of surface functional groups. The maximum NO3 adsorption capacity for the Fe-N-DF biochar at pH 3 was 19.4 mg g−1 (Table 5). The NO3 adsorption depends on pH, which is consistent with electrostatic attraction. Deprotonation of surface functional groups occurs as the pH increases. Moreover, OH ions become dominant adsorption sites with increasing pH and compete with NO3 [72]. NO3 can also form hydrogen bonds to protonated surface hydroxyls on the Fe oxides [44].
High-resolution XPS analysis was performed to understand biochar’s surface chemistry (See Figure 9). The peak N1s could be observed in the spectrum of N-DF and Fe-N-DF biochars. The N 1s spectra were deconvoluted into the following five peaks: pyridinic-N, pyridone-N, pyrrolic-N, graphitic-N, and oxidized-N. The pyridine functional group at 398.2 eV is the main nitrogen functional group in N-DF and Fe-N-DF biochars (Figure 9). Pyrrolic and pyridone groups on the biochar surface are neutral. However, pyridinic groups protonate at pH below pKa, resulting in +1 formal charge [34]. The pH of a solution is a crucial parameter influencing biochar surface charge because it can affect the protonation and deprotonation of surface functional groups, leading to surface charge change [73]. Protonation of biochar surface functional groups occurs at low solution pH, resulting in increased positive surface charge with the consequent impact on NO3 adsorption [74,75]. Pyridinic-N could contribute with one electron to the aromatic π-system and shows Lewis basic characteristics [76]. Previous studies showed that N doping altered the N-configuration on the char surface and caused a positive effect on the adsorption of negatively charged ions [28,35,47].
XRD results are shown in Figure 10. The diffraction peak at 2θ of 26.3° was assigned to the (002) diffraction of amorphous carbon [28]. The presence of α-Fe2O3, Fe3O4, and Fe3C can be observed in the XRD results. The formation of Fe3C resulted from the reduction of Fe oxide in the presence of N [77,78]. Diffraction peaks around 2θ of 40.7°, 42.8°, 44.6°, and 75.8° were attributed to crystalline phases of Fe3C. Fe3O4 can be formed during the pyrolysis of FeCl3-loaded DF under NH3 gas [79]. FeCl3 is hydrolyzed and converted to FeO(OH) during the drying of FeCl3-loaded biomass. Therefore, FeO/(OH) was converted to Fe3O4 by the reducing components H2 and CO. H2 is produced from NH3 at high temperatures, promoting the formation of Fe3O4 (Reactions (6)–(9)) [79]. The XRD results show no major changes in peak locations resulting from NO3 adsorption, suggesting that the adsorption process does not significantly alter the crystalline structure in biochar structure. This result also suggests that electrostatic attraction is the primary mechanism of NO3 adsorption.
FeCl3+ 3H2O → Fe(OH)3 + 3HCl
Fe(OH)3 → FeO(OH) + H2O
6FeO(OH) + 4H2 → 2 Fe3O4 + 4H2O
6FeO(OH) + 4CO → 2 Fe3O4 + 4CO2
The effects of pyrolysis temperature, metal loading, and pyrolysis residence time on the NO3 adsorption capacity of Fe-N-DF biochar were also investigated. For the experimental optimization strategies with a central composite experimental design for Fe-N-DF biochar, we varied the metal loading from 2 to 6 g per 15 g of Douglas fir wood, pyrolysis time between 30 and 90 min, and final pyrolysis temperature between 600 and 800 °C as input variables to identify the conditions resulting in a product with the highest NO3 adsorption capacity (see Figure 11).
The optimum pyrolysis temperature, metal loading, and pyrolysis time were 800 °C, 6 g, and 30 min, respectively. The NO3 adsorption capacity of Fe-N-DF biochar at pH 3 and 7 are 20.67 and 9.4 mg g−1, respectively, showing that the adsorption capacity of Fe-N-DF biochar (9.4 mg g−1) at a pH solution of 7 is higher by around 96% compared to pristine biochar (DF biochar: 4.8 mg g−1). The higher NO3 adsorption at pH 3 is attributed to the protonation of biochar surface functional groups. At low pH, protonation of biochar surface functional groups occurs, which causes an increasing positive surface charge [74,75]. Generally, the surface of raw biochar is negatively charged. Nevertheless, oxides of metals (e.g., Al2O3, MnO2, Ca2O3, Fe2O3, MgO2) and N-containing functional groups such as the pyridinic group in the biochar structure can make biochar more positively charged and, therefore, they can contribute to the adsorption of NO3 [80]. Table 6 compares this work’s experimental results with those reported in the literature for other materials. The results suggest that doping biochar with N and Fe greatly improves NO3 adsorption capacity from aqueous solutions. Therefore, co-doping biochar with Fe and N appears as a viable and easy-to-conduct strategy for improving the capacity of biochar to remove NO3 from water sources.

4. Conclusions

Novel Fe-N-co-doped biochar was produced via impregnation of Douglas fir with FeCl3 solution, followed by pyrolysis under NH3 gas. XRD, XPS, SEM, and EDS analyses revealed the presence of metal oxides and nitrogen-functional groups on the Fe-N-DF biochar. Although N-doped biochar exhibited a higher surface area (713 m2 g−1) than Fe-N-doped biochar (438 m2 g−1), the combined role of N and Fe functional groups in Fe-N-co-doped biochar is more important than large surface areas for NO3 adsorption. Therefore, the introduction of Fe and N in the biochar structure enhanced NO3 adsorption capacity. The presence of N-functional groups, especially pyridinic N, and Fe oxides on the synthesized Fe-N biochar provide more adsorption sites for NO3 adsorption, contributing to the high performance of NO3 removal from aqueous solution. The results show that decreasing the pH of the sorption medium caused improved NO3 adsorption capacity. The optimum process condition for producing Fe-N-DF biochar was found to be at a pyrolysis temperature of 800 °C, with a metal loading of 6 g per 15 g of biomass and a pyrolysis time of 30 min. These conditions resulted in a maximum NO3 adsorption capacity of 20.67 milligrams per gram at a pH of 3. Electrostatic attraction has been identified as the primary mechanism of NO3 adsorption.

Author Contributions

Conceptualization, S.H.M. and M.G.-P.; methodology, S.H.M.; experimental work, S.H.M. and K.M.; validation, S.H.M.; formal analysis, S.H.M. and Y.H.; investigation, S.H.M.; writing—original draft, S.H.M. and M.G.-P.; writing—review and editing, S.H.M., M.R.P.-S. and M.G.-P.; supervision, M.G.-P.; project administration, M.G.-P. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Sun Grant Initiative (Federal USDA/NIFA award AWD004255), and USDA/NIFA, through Hatch Project No. WNP00701.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in the study are included in the articlel, further inquiries can be directed to the corresponding author.

Acknowledgments

Special thanks to Jonathan Lamber, from the WSU Analytica Chemistry Service Center, for his assistance with the experimental work, and Oscar Marin-Flores for his support in the analysis of results.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Yin, Q.; Ren, H.; Wang, R.; Zhao, Z. Evaluation of nitrate and phosphate adsorption on Al-modified biochar: Influence of Al content. Sci. Total Environ. 2018, 631–632, 895–903. [Google Scholar] [CrossRef]
  2. Hornung, M. 11—The Role of Nitrates in the Eutrophication and Acidification of Surface Waters. In Managing Risks of Nitrates to Humans and the Environment; Wilson, W.S., Ball, A.S., Hinton, R.H., Eds.; Woodhead Publishing: Sawston, UK, 1999; pp. 155–174. [Google Scholar]
  3. Loganathan, P.; Vigneswaran, S.; Kandasamy, J. Enhanced removal of nitrate from water using surface modification of adsorbents—A review. J. Environ. Manag. 2013, 131, 363–374. [Google Scholar] [CrossRef]
  4. Haghighi Mood, S. Contribution to Nutrient Recovery from Anaerobically Digested Liquid Effluent Using Engineered Biochar; Washington State University: Pullman, WA, USA, 2022. [Google Scholar]
  5. Hwang, S.-J. Eutrophication and the Ecological Health Risk. Int. J. Environ. Res. Public Health 2020, 17, 6332. [Google Scholar] [CrossRef]
  6. Liu, Y.; Zhang, X.; Wang, J. A critical review of various adsorbents for selective removal of nitrate from water: Structure, performance and mechanism. Chemosphere 2022, 291, 132728. [Google Scholar] [CrossRef]
  7. Bhatnagar, A.; Sillanpää, M. A review of emerging adsorbents for nitrate removal from water. Chem. Eng. J. 2011, 168, 493–504. [Google Scholar] [CrossRef]
  8. Chiu, H.-F.; Tsai, S.-S.; Yang, C.-Y. Nitrate in Drinking Water and Risk of Death from Bladder Cancer: An Ecological Case-Control Study in Taiwan. J. Toxicol. Environ. Health Part A 2007, 70, 1000–1004. [Google Scholar] [CrossRef]
  9. Majumdar, D.; Gupta, N. Nitrate pollution of groundwater and associated human health disorders. Indian J. Environ. Health 2000, 42, 28–39. [Google Scholar]
  10. EPA. 2018 Edition of the Drinking Water Standards and Health Advisories Tables; Environmental Protection Agency: Washington, DC, USA, 2018. [Google Scholar]
  11. WHO. Nitrate and Nitrite in Drinking-Water Background Document for Development of WHO Guidelines for Drinking-Water Quality; WHO: Geneva, Switzerland, 2016. [Google Scholar]
  12. Adimalla, N.; Qian, H. Groundwater chemistry, distribution and potential health risk appraisal of nitrate enriched groundwater: A case study from the semi-urban region of South India. Ecotoxicol. Environ. Saf. 2021, 207, 111277. [Google Scholar] [CrossRef]
  13. Liu, J.; Peng, Y.; Li, C.; Gao, Z.; Chen, S. Characterization of the hydrochemistry of water resources of the Weibei Plain, Northern China, as well as an assessment of the risk of high groundwater nitrate levels to human health. Environ. Pollut. 2021, 268, 115947. [Google Scholar] [CrossRef]
  14. Abascal, E.; Gómez-Coma, L.; Ortiz, I.; Ortiz, A. Global diagnosis of nitrate pollution in groundwater and review of removal technologies. Sci. Total Environ. 2022, 810, 152233. [Google Scholar] [CrossRef]
  15. Pennino, M.J.; Compton, J.E.; Leibowitz, S.G. Trends in Drinking Water Nitrate Violations Across the United States. Environ. Sci. Technol. 2017, 51, 13450–13460. [Google Scholar] [CrossRef]
  16. Harter, T.; Lund, J.R.; Darby, J.; Fogg, G.E.; Howitt, R.; Jessoe, K.K.; Pettygrove, G.S.; Quinn, J.F.; Viers, J.H.; Boyle, D.B.; et al. Addressing Nitrate in California’s Drinking Water with a Focus on Tulare Lake Basin and Salinas Valley Groundwater. In State Water Resources Control Board Report to the Legislature; Center for Watershed Sciences, University of California: Davis, CA, USA, 2012; 78p. [Google Scholar]
  17. Divband Hafshejani, L.; Hooshmand, A.; Naseri, A.A.; Mohammadi, A.S.; Abbasi, F.; Bhatnagar, A. Removal of nitrate from aqueous solution by modified sugarcane bagasse biochar. Ecol. Eng. 2016, 95, 101–111. [Google Scholar] [CrossRef]
  18. Pang, Y.; Wang, J. Various electron donors for biological nitrate removal: A review. Sci. Total Environ. 2021, 794, 148699. [Google Scholar] [CrossRef]
  19. Rajta, A.; Bhatia, R.; Setia, H.; Pathania, P. Role of heterotrophic aerobic denitrifying bacteria in nitrate removal from wastewater. J. Appl. Microbiol. 2020, 128, 1261–1278. [Google Scholar] [CrossRef]
  20. Demiral, H.; Gündüzoğlu, G. Removal of nitrate from aqueous solutions by activated carbon prepared from sugar beet bagasse. Bioresour. Technol. 2010, 101, 1675–1680. [Google Scholar] [CrossRef]
  21. Priya, E.; Kumar, S.; Verma, C.; Sarkar, S.; Maji, P.K. A comprehensive review on technological advances of adsorption for removing nitrate and phosphate from waste water. J. Water Process Eng. 2022, 49, 103159. [Google Scholar] [CrossRef]
  22. Araújo, R.; Castro, A.C.M.; Santos Baptista, J.; Fiúza, A. Nanosized iron based permeable reactive barriers for nitrate removal—Systematic review. Phys. Chem. Earth Parts A/B/C 2016, 94, 29–34. [Google Scholar] [CrossRef]
  23. Fernández-López, J.A.; Alacid, M.; Obón, J.M.; Martínez-Vives, R.; Angosto, J.M. Nitrate-Polluted Waterbodies Remediation: Global Insights into Treatments for Compliance. Appl. Sci. 2023, 13, 4154. [Google Scholar] [CrossRef]
  24. Mazarji, M.; Aminzadeh, B.; Baghdadi, M.; Bhatnagar, A. Removal of nitrate from aqueous solution using modified granular activated carbon. J. Mol. Liq. 2017, 233, 139–148. [Google Scholar] [CrossRef]
  25. Haghighi Mood, S.; Pelaez-Samaniego, M.R.; Garcia-Perez, M. Perspectives of Engineered Biochar for Environmental Applications: A Review. Energy Fuels 2022, 36, 7940–7986. [Google Scholar] [CrossRef]
  26. Gęca, M.; Khalil, A.M.; Tang, M.; Bhakta, A.K.; Snoussi, Y.; Nowicki, P.; Wiśniewska, M.; Chehimi, M.M. Surface Treatment of Biochar&mdash;Methods, Surface Analysis and Potential Applications: A Comprehensive Review. Surfaces 2023, 6, 179–213. [Google Scholar] [CrossRef]
  27. Wang, W.; Zhu, Q.; Huang, R.; Hu, Y. Adsorption of nitrate in water by CTAB-modified MgFe layered double hydroxide composite biochar at low temperature: Adsorption characteristics and mechanisms. J. Environ. Chem. Eng. 2023, 11, 109090. [Google Scholar] [CrossRef]
  28. Haghighi Mood, S.; Ayiania, M.; Cao, H.; Marin-Flores, O.; Milan, Y.J.; Garcia-Perez, M. Nitrogen and magnesium Co-doped biochar for phosphate adsorption. Biomass Convers. Biorefinery 2021, 14, 5923–5942. [Google Scholar] [CrossRef]
  29. Quinn, K.; Mood, S.H.; Cervantes, E.; Perez, M.G.; Abu-Lail, N.I. Forces Governing the Transport of Pathogenic and Nonpathogenic Escherichia coli in Nitrogen and Magnesium Doped Biochar Amended Sand Columns. Microbiol. Res. 2023, 14, 218–228. [Google Scholar] [CrossRef]
  30. Yin, Q.; Zhang, B.; Wang, R.; Zhao, Z. Biochar as an adsorbent for inorganic nitrogen and phosphorus removal from water: A review. Environ. Sci. Pollut. Res. 2017, 24, 26297–26309. [Google Scholar] [CrossRef]
  31. Haghighi Mood, S.; Mainalis, K.; Pelaez-Samaniego, M.R.; Garcia-Perez, M. Characteristics of biochar Micro- and nano-chemical properties and interactions. In Biochar for Environmental Management, 3rd ed.; Lehmann, J., Joseph, S., Eds.; Routledge: London, UK, 2024; p. 25. [Google Scholar]
  32. Saha, D.; Thorpe, R.; Van Bramer, S.E.; Alexander, N.; Hensley, D.K.; Orkoulas, G.; Chen, J. Synthesis of Nitrogen and Sulfur Codoped Nanoporous Carbons from Algae: Role in CO2 Separation. ACS Omega 2018, 3, 18592–18602. [Google Scholar] [CrossRef]
  33. Boehm, H.P. Some aspects of the surface chemistry of carbon blacks and other carbons. Carbon 1994, 32, 759–769. [Google Scholar] [CrossRef]
  34. Lawrinenko, M.; Laird, D.A. Anion exchange capacity of biochar. Green Chem. 2015, 17, 4628–4636. [Google Scholar] [CrossRef]
  35. Mood, S.H.; Ayiania, M.; Jefferson-Milan, Y.; Garcia-Perez, M. Nitrogen doped char from anaerobically digested fiber for phosphate removal in aqueous solutions. Chemosphere 2020, 240, 124889. [Google Scholar] [CrossRef]
  36. Mainali, K.; Mood, S.H.; Pelaez-Samaniego, M.R.; Sierra-Jimenez, V.; Garcia-Perez, M. Production and applications of N-doped carbons from bioresources: A review. Catal. Today 2023, 423, 114248. [Google Scholar] [CrossRef]
  37. Ding, D.; Yang, S.; Qian, X.; Chen, L.; Cai, T. Nitrogen-doping positively whilst sulfur-doping negatively affect the catalytic activity of biochar for the degradation of organic contaminant. Appl. Catal. B Environ. 2020, 263, 118348. [Google Scholar] [CrossRef]
  38. Yoo, P.; Amano, Y.; Machida, M. Adsorption of nitrate onto nitrogen-doped activated carbon fibers prepared by chemical vapor deposition. Korean J. Chem. Eng. 2018, 35, 2468–2473. [Google Scholar] [CrossRef]
  39. Iida, T.; Amano, Y.; Machida, M.; Imazeki, F. Effect of surface property of activated carbon on adsorption of nitrate ion. Chem. Pharm. Bull. 2013, 61, 1173–1177. [Google Scholar] [CrossRef]
  40. Jellali, S.; El-Bassi, L.; Charabi, Y.; Usman, M.; Khiari, B.; Al-Wardy, M.; Jeguirim, M. Recent advancements on biochars enrichment with ammonium and nitrates from wastewaters: A critical review on benefits for environment and agriculture. J. Environ. Manag. 2022, 305, 114368. [Google Scholar] [CrossRef]
  41. Long, L.; Xue, Y.; Hu, X.; Zhu, Y. Study on the influence of surface potential on the nitrate adsorption capacity of metal modified biochar. Environ. Sci. Pollut. Res. 2019, 26, 3065–3074. [Google Scholar] [CrossRef]
  42. Min, L.; Zhongsheng, Z.; Zhe, L.; Haitao, W. Removal of nitrogen and phosphorus pollutants from water by FeCl3- impregnated biochar. Ecol. Eng. 2020, 149, 105792. [Google Scholar] [CrossRef]
  43. Ahmad, M.; Ahmad, M.; Usman, A.R.A.; Al-Faraj, A.S.; Abduljabbar, A.S.; Al-Wabel, M.I. Biochar composites with nano zerovalent iron and eggshell powder for nitrate removal from aqueous solution with coexisting chloride ions. Environ. Sci. Pollut. Res. 2018, 25, 25757–25771. [Google Scholar] [CrossRef]
  44. Bombuwala Dewage, N.; Liyanage, A.S.; Pittman, C.U.; Mohan, D.; Mlsna, T. Fast nitrate and fluoride adsorption and magnetic separation from water on α-Fe2O3 and Fe3O4 dispersed on Douglas fir biochar. Bioresour. Technol. 2018, 263, 258–265. [Google Scholar] [CrossRef]
  45. Dubinin, M.M.; Zaverina, E.D.; Radushkevich, L.V. Sorption and structure of active carbons. I. Adsorption of organic vapors. Zhurnal Fiz. Khimii 1947, 21, 1351–1362. [Google Scholar]
  46. Dubinin, M.M.; Radushkevich, L. Equation of the characteristic curve of activated charcoal. Chem. Zentr. 1947, 1, 875. [Google Scholar]
  47. Ayiania, M.; Garcia, A.; Haghighi Mood, S.; McEwen, J.-S.; Garcia-Perez, M. Novel Amorphous Carbons for the Adsorption of Phosphate: Part I. Elucidation of Chemical Structure of N-Metal-Doped Chars. ACS Omega 2022, 7, 14490–14504. [Google Scholar] [CrossRef]
  48. Hach. Chromotropic Acid Method, Method 10020; Hach: Ames, IA, USA, 2015. [Google Scholar]
  49. Li, D.-C.; Ding, J.-W.; Qian, T.-T.; Zhang, S.; Jiang, H. Preparation of high adsorption performance and stable biochar granules by FeCl3-catalyzed fast pyrolysis. RSC Adv. 2016, 6, 12226–12234. [Google Scholar] [CrossRef]
  50. Wan, Z.; Sun, Y.; Tsang, D.C.W.; Khan, E.; Yip, A.C.K.; Ng, Y.H.; Rinklebe, J.; Ok, Y.S. Customised fabrication of nitrogen-doped biochar for environmental and energy applications. Chem. Eng. J. 2020, 401, 126136. [Google Scholar] [CrossRef]
  51. Luo, W.; Wang, B.; Heron, C.G.; Allen, M.J.; Morre, J.; Maier, C.S.; Stickle, W.F.; Ji, X. Pyrolysis of Cellulose under Ammonia Leads to Nitrogen-Doped Nanoporous Carbon Generated through Methane Formation. Nano Lett. 2014, 14, 2225–2229. [Google Scholar] [CrossRef]
  52. Khan, Z.H.; Gao, M.; Qiu, W.; Islam, M.S.; Song, Z. Mechanisms for cadmium adsorption by magnetic biochar composites in an aqueous solution. Chemosphere 2020, 246, 125701. [Google Scholar] [CrossRef]
  53. Yang, J.; Zhao, Y.; Ma, S.; Zhu, B.; Zhang, J.; Zheng, C. Mercury Removal by Magnetic Biochar Derived from Simultaneous Activation and Magnetization of Sawdust. Environ. Sci. Technol. 2016, 50, 12040–12047. [Google Scholar] [CrossRef]
  54. Zhao, J.; Xu, H.; Tang, D.; Mathews, J.P.; Li, S.; Tao, S. A comparative evaluation of coal specific surface area by CO2 and N2 adsorption and its influence on CH4 adsorption capacity at different pore sizes. Fuel 2016, 183, 420–431. [Google Scholar] [CrossRef]
  55. Jagiello, J.; Ania, C.; Parra, J.B.; Cook, C. Dual gas analysis of microporous carbons using 2D-NLDFT heterogeneous surface model and combined adsorption data of N2 and CO2. Carbon 2015, 91, 330–337. [Google Scholar] [CrossRef]
  56. Wang, Y.; Guo, W.; Chen, W.; Xu, G.; Zhu, G.; Xie, G.; Xu, L.; Dong, C.; Gao, S.; Chen, Y.; et al. Co-production of porous N-doped biochar and hydrogen-rich gas production from simultaneous pyrolysis-activation-nitrogen doping of biomass: Synergistic mechanism of KOH and NH3. Renew. Energy 2024, 229, 120777. [Google Scholar] [CrossRef]
  57. Li, D.; Chen, W.; Wu, J.; Jia, C.Q.; Jiang, X. The preparation of waste biomass-derived N-doped carbons and their application in acid gas removal: Focus on N functional groups. J. Mater. Chem. A 2020, 8, 24977–24995. [Google Scholar] [CrossRef]
  58. Wang, X.J.; Wang, Y.; Wang, X.; Liu, M.; Xia, S.Q.; Yin, D.Q.; Zhang, Y.L.; Zhao, J.F. Microwave-assisted preparation of bamboo charcoal-based iron-containing adsorbents for Cr(VI) removal. Chem. Eng. J. 2011, 174, 326–332. [Google Scholar] [CrossRef]
  59. Zhu, J.; Gu, H.; Guo, J.; Chen, M.; Wei, H.; Luo, Z.; Colorado, H.A.; Yerra, N.; Ding, D.; Ho, T.C.; et al. Mesoporous magnetic carbon nanocomposite fabrics for highly efficient Cr(vi) removal. J. Mater. Chem. A 2014, 2, 2256–2265. [Google Scholar] [CrossRef]
  60. Toncón-Leal, C.F.; Villarroel-Rocha, J.; Silva, M.T.P.; Braga, T.P.; Sapag, K. Characterization of mesoporous region by the scanning of the hysteresis loop in adsorption–desorption isotherms. Adsorption 2021, 27, 1109–1122. [Google Scholar] [CrossRef]
  61. Bi, F.; Feng, X.; Zhou, Z.; Zhang, Y.; Wei, J.; Yuan, L.; Liu, B.; Huang, Y.; Zhang, X. Mn-based catalysts derived from the non-thermal treatment of Mn-MIL-100 to enhance its water-resistance for toluene oxidation: Mechanism study. Chem. Eng. J. 2024, 485, 149776. [Google Scholar] [CrossRef]
  62. Feng, C.; Chen, C.; Wang, J.; Li, S.; Liu, F.; Pan, Y.; Lu, Y.; Liu, Y.; Li, X.; Liu, Y.; et al. Insights into the synergistic effect and catalytic mechanism in MnCeOx solid solution catalysts for low-temperature propane total oxidation. Surf. Interfaces 2024, 46, 104011. [Google Scholar] [CrossRef]
  63. Schneider, P. Adsorption isotherms of microporous-mesoporous solids revisited. Appl. Catal. A Gen. 1995, 129, 157–165. [Google Scholar] [CrossRef]
  64. Jiang, S.; Yan, L.; Wang, R.; Li, G.; Rao, P.; Ju, M.; Jian, L.; Guo, X.; Che, L. Recyclable nitrogen-doped biochar via low-temperature pyrolysis for enhanced lead(II) removal. Chemosphere 2022, 286, 131666. [Google Scholar] [CrossRef]
  65. Oh, W.-D.; Veksha, A.; Chen, X.; Adnan, R.; Lim, J.-W.; Leong, K.-H.; Lim, T.-T. Catalytically active nitrogen-doped porous carbon derived from biowastes for organics removal via peroxymonosulfate activation. Chem. Eng. J. 2019, 374, 947–957. [Google Scholar] [CrossRef]
  66. Alsewailem, F.D. Modified Correlations for Adsorption Isotherms. J. Chem. Eng. Data 2015, 60, 762–765. [Google Scholar] [CrossRef]
  67. Latour, R.A. The Langmuir isotherm: A commonly applied but misleading approach for the analysis of protein adsorption behavior. J. Biomed. Mater. Res. Part A 2015, 103, 949–958. [Google Scholar] [CrossRef]
  68. Ghosal, P.S.; Gupta, A.K. Development of a generalized adsorption isotherm model at solid-liquid interface: A novel approach. J. Mol. Liq. 2017, 240, 21–24. [Google Scholar] [CrossRef]
  69. Umpleby, R.J.; Baxter, S.C.; Chen, Y.; Shah, R.N.; Shimizu, K.D. Characterization of Molecularly Imprinted Polymers with the Langmuir−Freundlich Isotherm. Anal. Chem. 2001, 73, 4584–4591. [Google Scholar] [CrossRef] [PubMed]
  70. Saadi, R.; Saadi, Z.; Fazaeli, R.; Fard, N.E. Monolayer and multilayer adsorption isotherm models for sorption from aqueous media. Korean J. Chem. Eng. 2015, 32, 787–799. [Google Scholar] [CrossRef]
  71. Foo, K.Y.; Hameed, B.H. Insights into the modeling of adsorption isotherm systems. Chem. Eng. J. 2010, 156, 2–10. [Google Scholar] [CrossRef]
  72. Khoshkalam, E.; Fotovat, A.; Halajnia, A.; Kazemian, H.; Eshghi, H. Nitrate adsorption using green iron oxide nanoparticles synthesized by Eucalyptus leaf extracts: Kinetics and effects of pH, KCl salt, and anions competition. J. Mol. Liq. 2023, 375, 121366. [Google Scholar] [CrossRef]
  73. Liu, J.; Jiang, J.; Meng, Y.; Aihemaiti, A.; Xu, Y.; Xiang, H.; Gao, Y.; Chen, X. Preparation, environmental application and prospect of biochar-supported metal nanoparticles: A review. J. Hazard. Mater. 2020, 388, 122026. [Google Scholar] [CrossRef]
  74. Li, R.; Wang, J.J.; Zhou, B.; Awasthi, M.K.; Ali, A.; Zhang, Z.; Gaston, L.A.; Lahori, A.H.; Mahar, A. Enhancing phosphate adsorption by Mg/Al layered double hydroxide functionalized biochar with different Mg/Al ratios. Sci. Total Environ. 2016, 559, 121–129. [Google Scholar] [CrossRef] [PubMed]
  75. Li, W.; Pierre-Louis, A.-M.; Kwon, K.D.; Kubicki, J.D.; Strongin, D.R.; Phillips, B.L. Molecular level investigations of phosphate sorption on corundum (α-Al2O3) by 31P solid state NMR, ATR-FTIR and quantum chemical calculation. Geochim. Cosmochim. Acta 2013, 107, 252–266. [Google Scholar] [CrossRef]
  76. Petrovic, B.; Gorbounov, M.; Masoudi Soltani, S. Impact of Surface Functional Groups and Their Introduction Methods on the Mechanisms of CO2 Adsorption on Porous Carbonaceous Adsorbents. Carbon Capture Sci. Technol. 2022, 3, 100045. [Google Scholar] [CrossRef]
  77. Mei, Y.; Xu, J.; Zhang, Y.; Li, B.; Fan, S.; Xu, H. Effect of Fe–N modification on the properties of biochars and their adsorption behavior on tetracycline removal from aqueous solution. Bioresour. Technol. 2021, 325, 124732. [Google Scholar] [CrossRef]
  78. Yan, X.-H.; Xu, B.-Q. Mesoporous carbon material co-doped with nitrogen and iron (Fe–N–C): High-performance cathode catalyst for oxygen reduction reaction in alkaline electrolyte. J. Mater. Chem. A 2014, 2, 8617–8622. [Google Scholar] [CrossRef]
  79. Mian, M.M.; Liu, G.; Yousaf, B.; Fu, B.; Ullah, H.; Ali, M.U.; Abbas, Q.; Mujtaba Munir, M.A.; Ruijia, L. Simultaneous functionalization and magnetization of biochar via NH3 ambiance pyrolysis for efficient removal of Cr (VI). Chemosphere 2018, 208, 712–721. [Google Scholar] [CrossRef] [PubMed]
  80. Alkurdi, S.S.A.; Herath, I.; Bundschuh, J.; Al-Juboori, R.A.; Vithanage, M.; Mohan, D. Biochar versus bone char for a sustainable inorganic arsenic mitigation in water: What needs to be done in future research? Environ. Int. 2019, 127, 52–69. [Google Scholar] [CrossRef]
  81. Wang, Z.; Guo, H.; Shen, F.; Yang, G.; Zhang, Y.; Zeng, Y.; Wang, L.; Xiao, H.; Deng, S. Biochar produced from oak sawdust by Lanthanum (La)-involved pyrolysis for adsorption of ammonium (NH4+), nitrate (NO3), and phosphate (PO43−). Chemosphere 2015, 119, 646–653. [Google Scholar] [CrossRef]
  82. Zhao, H.; Xue, Y.; Long, L.; Hu, X. Adsorption of nitrate onto biochar derived from agricultural residuals. Water Sci. Technol. 2017, 77, 548–554. [Google Scholar] [CrossRef] [PubMed]
  83. Viglašová, E.; Galamboš, M.; Danková, Z.; Krivosudský, L.; Lengauer, C.L.; Hood-Nowotny, R.; Soja, G.; Rompel, A.; Matík, M.; Briančin, J. Production, characterization and adsorption studies of bamboo-based biochar/montmorillonite composite for nitrate removal. Waste Manag. 2018, 79, 385–394. [Google Scholar] [CrossRef]
  84. Vijayaraghavan, K.; Balasubramanian, R. Application of pinewood waste-derived biochar for the removal of nitrate and phosphate from single and binary solutions. Chemosphere 2021, 278, 130361. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of the preparation processes of biochars studied.
Figure 1. Schematic representation of the preparation processes of biochars studied.
Sustainability 16 05733 g001
Figure 2. Pore size distribution of chars obtained from (a) CO2 and (b) N2 adsorption.
Figure 2. Pore size distribution of chars obtained from (a) CO2 and (b) N2 adsorption.
Sustainability 16 05733 g002
Figure 3. N2 adsorption isotherms for (a) DF biochar, (b) N-DF biochar, (c) Fe-DF biochar, and (d) Fe-N-DF biochar (biochars were produced at 800 °C for 60 min).
Figure 3. N2 adsorption isotherms for (a) DF biochar, (b) N-DF biochar, (c) Fe-DF biochar, and (d) Fe-N-DF biochar (biochars were produced at 800 °C for 60 min).
Sustainability 16 05733 g003
Figure 4. CO2 adsorption isotherms for (a) DF biochar, (b) N-DF biochar, (c) Fe-DF biochar, and (d) Fe-N-DF biochar.
Figure 4. CO2 adsorption isotherms for (a) DF biochar, (b) N-DF biochar, (c) Fe-DF biochar, and (d) Fe-N-DF biochar.
Sustainability 16 05733 g004
Figure 5. Three-dimensional response surface for surface area: (a) pyrolysis time vs. pyrolysis temperature, (b) metal loading vs. pyrolysis temperature, and (c) pyrolysis time vs. metal loading.
Figure 5. Three-dimensional response surface for surface area: (a) pyrolysis time vs. pyrolysis temperature, (b) metal loading vs. pyrolysis temperature, and (c) pyrolysis time vs. metal loading.
Sustainability 16 05733 g005
Figure 6. FTIR spectra of the four types of DF biochar, N-DF biochar, Fe-DF biochar, and Fe-N-DF biochar (produced at 800 °C for 60 min).
Figure 6. FTIR spectra of the four types of DF biochar, N-DF biochar, Fe-DF biochar, and Fe-N-DF biochar (produced at 800 °C for 60 min).
Sustainability 16 05733 g006
Figure 7. SEM and TEM images of Fe-N-DF biochar: (a) SEM at 40000X and (b) SEM at 4000X, (c) SEM image and EDS spectra from Fe-N-DF biochar, (d) TEM image of Fe-N-DF biochar, (e) TEM image of Fe-N-DF biochar; and (f) SEM image and EDS spectra from Fe-N-DF biochar.
Figure 7. SEM and TEM images of Fe-N-DF biochar: (a) SEM at 40000X and (b) SEM at 4000X, (c) SEM image and EDS spectra from Fe-N-DF biochar, (d) TEM image of Fe-N-DF biochar, (e) TEM image of Fe-N-DF biochar; and (f) SEM image and EDS spectra from Fe-N-DF biochar.
Sustainability 16 05733 g007
Figure 8. Adsorption isotherm data (red dots) and modeling (lines) for NO3 on biochars (produced at 800 °C) at a solution pH of 7. A similar y-axis scale is used for easy comparison of results.
Figure 8. Adsorption isotherm data (red dots) and modeling (lines) for NO3 on biochars (produced at 800 °C) at a solution pH of 7. A similar y-axis scale is used for easy comparison of results.
Sustainability 16 05733 g008
Figure 9. N 1s XPS spectra of (a) N-DF and (b) Fe-N-DF biochars (PI—pyridinic group; PII—pyridone group; PIII—pyrrolic group; PIV—graphitic N; and PV—N-oxide group). Red, black, and green lines represent the experimental data, fitting results, and background, respectively. The small peaks created by the dotted lines represent the components identified through deconvolution.
Figure 9. N 1s XPS spectra of (a) N-DF and (b) Fe-N-DF biochars (PI—pyridinic group; PII—pyridone group; PIII—pyrrolic group; PIV—graphitic N; and PV—N-oxide group). Red, black, and green lines represent the experimental data, fitting results, and background, respectively. The small peaks created by the dotted lines represent the components identified through deconvolution.
Sustainability 16 05733 g009
Figure 10. XRD spectrum of the Fe-N-DF biochar (a) before and (b) after NO3 adsorption.
Figure 10. XRD spectrum of the Fe-N-DF biochar (a) before and (b) after NO3 adsorption.
Sustainability 16 05733 g010
Figure 11. Three-dimensional response surface for adsorption capacity. (a) Pyrolysis temperature versus metal loading; (b) pyrolysis temperature versus pyrolysis time; and (c) pyrolysis time versus metal loading.
Figure 11. Three-dimensional response surface for adsorption capacity. (a) Pyrolysis temperature versus metal loading; (b) pyrolysis temperature versus pyrolysis time; and (c) pyrolysis time versus metal loading.
Sustainability 16 05733 g011
Table 1. Independent factors and their coded levels for the CCD.
Table 1. Independent factors and their coded levels for the CCD.
FactorsCodesCoded Factors’ Level
Low (−1)Center (0)High (+1)
Pyrolysis temperature (°C)A600700800
Pyrolysis time (min)B306090
Metal loading * (g/per 15 g of DF)C246
* Metal loading is defined as the amount of FeCl3 in grams per 15 g of biomass in 200 mL of water.
Table 2. Results on the proximate analysis of the four types of biochar.
Table 2. Results on the proximate analysis of the four types of biochar.
SampleVolatile MatterFixed CarbonAsh
DF biochar7.23 ± 0.1892.23 ± 0.650.53 ± 0.47
N-DF biochar7.06 ± 0.2192.7 ± 0.250.24 ± 0.04
Fe-DF biochar8.07 ± 0.0289.01 ± 1.082.92 ± 0.07
Fe-N-DF biochar13.40 ± 0.4783.24 ± 0.623.37 ± 0.01
Table 3. Elemental composition (wt. %) and C/N and C/O ratios of the biochars produced.
Table 3. Elemental composition (wt. %) and C/N and C/O ratios of the biochars produced.
SampleCHNO *C/OC/N
DF biochar93.62 ± 0.20.51 ± 0.040.60 ± 0.054.7419.77156.07
N-DF biochar80.09 ± 0.10.93 ± 0.036.65 ± 0.0512.086.6312.05
Fe-DF biochar91.91 ± 3.30.6 ± 0.020.61 ± 0.13.9623.18151.21
Fe-N-DF biochar80.73 ± 1.20.85 ± 0.058.03 ± 0.17.0211.510.05
* By difference.
Table 4. Surface areas and pore size volumes as determined by N2 and CO2 adsorptions.
Table 4. Surface areas and pore size volumes as determined by N2 and CO2 adsorptions.
SampleSaN2 (m2 g−1)SaCO2 (m2 g−1)Micro Volume (cm3 g−1)
DF Biochar478 ± 9.85330.21
N-DF Biochar713 ± 14.66940.28
Fe-DF Biochar431 ± 54940.20
Fe-N-DF Biochar438 ± 4.14770.19
Table 5. Isotherm parameters for adsorption of NO3.
Table 5. Isotherm parameters for adsorption of NO3.
BiocharpHLangmuirFreundlich
K
(L mg−1)
Q
(mg g−1)
R2K
(mg(1−n) Ln g−1)
nR2
DF30.2817.00.986.060.230.86
Fe-DF0.3816.10.996.380.230.88
N-DF0.4018.90.977.260.240.87
Fe-N-DF0.8319.40.979.010.200.85
DF70.124.80.881.470.250.81
Fe-DF0.116.30.971.690.280.98
N-DF0.158.70.902.470.280.95
Fe-N-DF1.449.370.975.170.150.86
DF80.093.50.880.940.270.87
Fe-DF0.125.00.961.640.230.80
N-DF0.396.30.832.940.180.90
Fe-N-DF0.537.30.923.490.180.99
BiocharpHLangmuir–FreundlichRedlich–Peterson
K
(Ln mg−n)
Q
(mg g−1)
nR2K
(L g−1)
a
(Ln mg−n)
nR2
DF30.2316.11.30.994.370.221.040.98
Fe-DF0.3816.11.00.996.340.410.990.99
N-DF0.3818.41.20.987.780.430.990.98
Fe-N-DF0.8918.91.20.9817.10.940.980.98
DF70.074.51.30.890.570.111.000.89
Fe-DF0.158.40.60.991.280.410.840.99
N-DF0.1319.40.40.954.131.220.790.96
Fe-N-DF1.099.80.70.9814.831.730.980.97
DF80.144.10.70.900.440.210.900.89
Fe-DF0.044.61.50.980.400.041.170.99
N-DF0.3111.30.30.919.062.630.860.91
Fe-N-DF0.3412.80.30.9912.412.990.860.99
Table 6. Overview of results reported in the literature on NO3 removal using different types of biochar.
Table 6. Overview of results reported in the literature on NO3 removal using different types of biochar.
FeedstockTreatment MethodSurface Area (m2 g−1)pH of the SolutionAdsorption Capacity (mg.g−1)References
Oak sawdustFeedstock impregnated
with LaCl2
23.08-8.7[81]
Oak sawdust-24.94-2.81[81]
Sugarcane bagasseChemical modification using epichlorohydrin, N, N-dimethylformamide, ethylenediamine and trimethylamine 41.674.6428.21[17]
Corncob---14.46[82]
Bamboo-2845[83]
BambooFeedstock impregnated
with Clay
15649[83]
Pinewood waste-204.224.2[84]
Douglas fir-47874.8This study
Douglas firAmmonia treatment71378.7This study
Douglas firFeedstock impregnated
with FeCl3
43176.3This study
Douglas firFeedstock was impregnated with FeCl3 and pyrolyzed under ammonia gas43879.4This study
Douglas firFeedstock was impregnated with FeCl3 and pyrolyzed under ammonia gas438320.67This study
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Haghighi Mood, S.; Pelaez-Samaniego, M.R.; Han, Y.; Mainali, K.; Garcia-Perez, M. Iron- and Nitrogen-Modified Biochar for Nitrate Adsorption from Aqueous Solution. Sustainability 2024, 16, 5733. https://doi.org/10.3390/su16135733

AMA Style

Haghighi Mood S, Pelaez-Samaniego MR, Han Y, Mainali K, Garcia-Perez M. Iron- and Nitrogen-Modified Biochar for Nitrate Adsorption from Aqueous Solution. Sustainability. 2024; 16(13):5733. https://doi.org/10.3390/su16135733

Chicago/Turabian Style

Haghighi Mood, Sohrab, Manuel Raul Pelaez-Samaniego, Yinglei Han, Kalidas Mainali, and Manuel Garcia-Perez. 2024. "Iron- and Nitrogen-Modified Biochar for Nitrate Adsorption from Aqueous Solution" Sustainability 16, no. 13: 5733. https://doi.org/10.3390/su16135733

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