Next Article in Journal
Progress in Ski Jumping Technology Based on Biomechanical Sport Research Methods
Next Article in Special Issue
Study on the Alkali–Sulfur Co-Activation and Mechanical Properties of Low-Carbon Cementitious Composite Materials Based on Electrolytic Manganese Residue, Carbide Slag, and Granulated Blast-Furnace Slag
Previous Article in Journal
Heterogeneous Graph-Convolution-Network-Based Short-Text Classification
Previous Article in Special Issue
Solidification of Radioactive Wastes Using Recycled Cement Originating from Decommissioned Nuclear-Energy Facilities
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Adsorption of Ammonium, Nitrate, and Phosphate on Hydrochars and Biochars

by
Paulo André Trazzi
1,2,
Mayank Vashishtha
2,
Jan Najser
3,
Achim Schmalenberger
4,
Vasanth Kumar Kannuchamy
5,
James J. Leahy
2 and
Witold Kwapinski
2,*
1
PPG Ciflor, Universidade Federal do Acre, UFAC, BR 364 km 04, Bairro Distrito Industrial, Rio Branco 69920-900, Brazil
2
Chemical Sciences Department, University of Limerick, V94 T9PX Limerick, Ireland
3
ENET Centre, VSB—Technical University of Ostrava, 708 00 Ostrava, Czech Republic
4
Biological Sciences Department, University of Limerick, V94 T9PX Limerick, Ireland
5
School of Chemistry and Chemical Engineering, Faculty of Engineering and Physical Sciences, University of Surrey, Guildford GU2 7XH, UK
*
Author to whom correspondence should be addressed.
Appl. Sci. 2024, 14(6), 2280; https://doi.org/10.3390/app14062280
Submission received: 5 December 2023 / Revised: 2 March 2024 / Accepted: 5 March 2024 / Published: 8 March 2024
(This article belongs to the Special Issue Sustainable Materials and Waste Recovery)

Abstract

:
Biochar (BC) and hydrochar (HC) have attracted considerable attention owing to their versatile characteristics and proven effectiveness in diverse technical fields. Solid BC is generated as a result of the dry carbonisation process of pyrolysis, in contrast to the slurry HC, which is produced during the hydrothermal carbonisation process. In this study, we evaluated the adsorption potential of two hydrochar samples (HCs) and three biochar samples (BCs) produced from sugar cane bagasse. The adsorption capacity of these samples was tested for ammonium, nitrate, and phosphate ions under various conditions. The BCs and HCs were subjected to characterisation using a CHNS/O analyser, the zeta potential, and Fourier transform infrared (FTIR). Elevating the pyrolysis temperature of the biochar resulted in changes in the fixed carbon and ash contents, while the volatile matter and H/C and O/C atomic ratios decreased. As the residence time increased, the H/C ratio and volatile matter content of the hydrochars (HCs) decreased. However, the fixed carbon content, ash content, and O/C and C/N ratios exhibited an increase. Thermodynamics, adsorption isotherms, and pH were also taken into consideration. The FTIR spectra analysis indicated that the carboxyl and ester functional groups present in both the BCs and HCs displayed reduced peak intensities subsequent to the adsorption of the three ions. While the adsorption was exothermic, we noticed that the adsorption capacity increased with temperature. The results indicate that sorption was homogenous across all binding sites, as evidenced by the optimal fit to the Langmuir isotherm. The research findings indicate that the adsorption capacity of various BC and HC adsorbents is significantly influenced by the surface area of the adsorbents in the case of nitrate and phosphate, but in the case of ammonia, adsorption is dictated by the functional polar groups present on the adsorbent surface.

1. Introduction

The nomenclature employed to refer to the residual material derived from the slow pyrolysis of biowaste, which is primarily composed of carbon, is biochar (BC). Conversely, the residual material obtained from the hydrothermal carbonisation (HTC) process, also called wet pyrolysis, is referred to as hydrochar (HC). Both of the aforementioned products have garnered significant attention owing to their potential applications in a variety of engineering and industrial fields [1]. The production of BC involves a pyrolysis process that necessitates an extra energy input to dry the feedstock. Conversely, the production of HC enables the utilisation of wet feedstock that does not require drying during the process. The biomass undergoes a high-temperature treatment in the presence of water, which serves as a catalyst [2].
BC and HC exhibit promising characteristics and structures that make them suitable for utilisation as soil enhancers. These features include the capacity to enhance crop yield, facilitate the filtration of percolating soil water, and assimilate atmospheric carbon in soil [3]. The variances in the characteristics and functions of BC that have been observed are subject to the influence of various process parameters [4,5,6,7,8]. These parameters include the feedstock, processing temperature, heating rate, and residence time. The synthesis of HC can be achieved through the utilisation of water as the reaction medium at a lower temperature, resulting in a more cost-effective process compared to pyrolysis. The reaction conditions commonly observed in the literature include temperatures ranging from 180 to 260 °C, autogenetic pressures ranging from 2 to 6 MPa, and reaction times ranging from 30 to 240 min [9]. The surface of HC exhibits a higher concentration of functional groups containing oxygen and molecules that have undergone aromatisation. The incorporation of functional groups into the molecule enhances its hydrophilicity, thereby rendering it suitable for augmenting the water retention capacity of soil [10].
Distinguishable variations exist in the physical and chemical characteristics of BCs and HCs. The study by Takaya and colleagues [11] revealed that BC and HC exhibit adsorption capacities ranging from around 0 to 30 mg·g−1 and 105 to 146 mg·g−1 for phosphate and ammonium, respectively. This observed phenomenon can be attributed primarily to the unique physicochemical characteristics of the subject in question. The investigation of the properties and interaction mechanisms of BC and HC adsorption is crucial for the utilisation of soil fertilisers in order to mitigate potential environmental consequences. Eutrophication is a phenomenon that occurs in estuaries and coastal waters as a result of an overabundance of nitrogen and phosphate. This process leads to an escalation in the growth of plant and toxic algae [12]. The application of synthetic fertilisers has been found to have a positive correlation with the release of ammonium, nitrate, and phosphorus into water bodies from agricultural land [13,14]. Research [15,16] findings indicate that BC and HC have the ability to impede the uptake of nitrogen and phosphorus by the soil. The utilisation of sugarcane bagasse BC has been observed to enhance the retention time of nitrate in the vicinity of plant roots, thereby facilitating its uptake by the plants [17]. The optimal conditions were identified for ammonium adsorption on bamboo charcoal, which resulted in the removal of up to 82.4% of ammonium from an aqueous solution [18]. As per existing research, it has been observed that BC has the ability to solubilise phosphate. It has also been noted that modifying the concentration of BC present in soil can lead to improved agronomic performance [19]. As per the findings of a previous study [20], it can be inferred that HC exhibits potential as an adsorbent for the purpose of removing and recovering phosphorus from wastewater. This is supported by the fact that HC has shown to be remarkably efficient in adsorbing phosphate ions under different pH conditions, even in the presence of competing anions.
The investigation of the sorption characteristics of BC and HC is currently in its nascent phase. The majority of research investigations have concentrated on the ion adsorption capabilities of BC, whereas the published findings regarding HC have been notably limited. Understanding the characteristics of both BCs and HCs and their impact on the adsorption mechanisms of ammonium, nitrate, and phosphate is crucial for the improved management of nitrogen and phosphorus in soil and agronomic practices. The aim of this investigation is to examine the impact of bio-based feedstock-derived BCs and HCs, produced under various dry and hydrothermal carbonisation conditions, on the adsorption of ammonium, nitrate, and phosphate.

2. Materials and Methods

2.1. BC and HC Preparation

BC and HC were obtained through the pyrolysis and thermal hydrolysis of sugar cane bagasse, respectively. The biomass underwent an initial shredding process resulting in fragments measuring between 2 and 3 cm in length. The study involved the implementation of slow pyrolysis at 300 °C, 500 °C, and 700 °C for a duration of 60 min within a fixed-bed reactor. The BC produced at these temperatures was labelled as BC300, BC500, and BC700, respectively. The experimental setup comprised a temperature controller cabinet, a quartz tube reactor with an inner diameter of 2.5 cm, and an electric furnace heater. The BCs that were acquired underwent a grinding process and were subsequently rinsed with deionised water. Afterward, they were dried in a furnace at a temperature of 105 °C for a duration of 24 h. The BCs were then stored in airtight containers at room temperature until they were ready for further use.
HC was produced by means of a 2G Parr agitated pressure reactor featuring a removable glass liner. The experimental procedure involved the addition of 200 g of sugar cane bagasse and 1 L of distilled water into the glass liner, followed by thorough mixing for a duration of 30 min. The vessel was hermetically sealed and subjected to a continuous flow of nitrogen gas for a duration of 10 min. The experiment involved subjecting the vessel and its contents to a temperature of 250 °C while simultaneously agitating them for 30 and 180 min. This process yielded two distinct samples, namely, HC30 and HC180. Subsequently, the vessel along with its contents underwent rapid cooling in a cold-water bath maintained at room temperature. The HCs were isolated through the process of vacuum filtration, followed by multiple washes with distilled water. The resulting HCs were then subjected to drying in an oven set at 105 °C for a period of 24 h. Finally, the HCs were stored in plastic receptacles for further use.
Activated carbon (AC) from Sigma-Aldrich was subjected to characterisation and subsequently employed in sorption experiments to facilitate a comparison between the BCs and HCs.

2.2. Characterisation of BCs and HCs

The BC, HC, and AC samples were subjected to analysis utilising a CHNS/O analyser, specifically the Elementar Variant LE Cube. The moisture content of the char samples was determined in accordance with the ICS 75.160.10, DD CEN/TS 14774-3:2009 standard [21]. The ash and volatile matter contents were determined using ICS 75.160.10, DD CEN/TS 14775:2009 [22] and ICS 75.160.10, DD CEN/TS 15148:2009 [23], respectively. The pH values of the BCs and HCs were measured by means of a pH metre (PHM 84, Radiometer, Denmark) equipped with glass REF 451 and calomel pHG 201-8 electrodes. The ratio of BC or HC to distilled water was maintained at 1:20 (m/v) during the pH determination process. The study employed a Fourier transform infrared (FTIR) spectrometer, specifically the Cary 630 FTIR spectrometer from Agilent Technologies Inc. (Santa Clara, CA, USA). Spectra were obtained within the 400–4000 cm−1 range, using a resolution of 4 cm−1 and 64 scans per sample. The biochar surface area and pore volume were determined by using the Brunauer–Emmett–Teller (BET) method, using a Gemini 2375 V5.01 surface area analyser (Mcromeritics Instrument Co., Norcross, GA, USA).
The zeta potentials of the BCs, HCs, and AC were determined. The experimental procedure involved mixing 0.5 g of each sample with 100 mL of deionised water at a pH of 7. The impact of NH4+, NO3, and PO43− adsorption on the zeta potential was assessed by introducing 0.5 g of each char sample to a solution containing 100 mg·dm−3 of NH4+, NO3, and PO43−. Prior to this, the pH of the suspensions was adjusted to a range between 4.0 and 9.0 using 0.1 mol·dm−3 of HCl or NaOH. The solutions under investigation were subjected to mechanical agitation at a rate of 250 revolutions per minute for a duration of 30 min. The suspensions underwent ultrasonic dispersion for a duration of 1 h at a temperature of 30 °C, utilising a bath-type sonicator with a frequency of 40 kHz and a voltage of 300 W. Subsequently, the solution underwent filtration using filter paper of the Whatman 42 variety. The zeta potential of each supernatant solution was determined by means of a Malvern Zetasizer Nano instrument (Malvern Instruments, Malvern, UK). The replicability of all analyses was ensured.

2.3. Adsorption of NH4+, NO3, and PO43−

Initial analyses were performed to determine the most effective adsorber dosage. Adsorption experiments were conducted over a range of adsorbent masses, specifically between 0.1 and 1 g with increments of 0.1 g. The concentration of the adsorbate used was 100 mg·dm−3. A notable variance in the adsorber material dosage was observed at 0.5 g, which was utilised in subsequent experimental trials.

2.3.1. Analysis of NH4+, NO3, and PO43−

An analysis of NH4+ was performed utilising a Varian Spectrophotometer (UV-Vis-4000) in accordance with the phenate method [24]. The determination of NO3 and PO43− was carried out using ion chromatography equipment (Dionex DX-120 Ion and ICS1100, Manasquan, NJ, USA). The analysis points were obtained by calculating the mean of three distinct parallel sample solutions. Statistical significance was determined at a level of p < 0.05.

2.3.2. pH Effect

In order to determine the effect of pH, 0.5 g of an adsorbent material was mixed with 100 cm3 of a solution containing 100 mg·dm−3 of either NH4+, NO3, or PO43−. Before adding the biochar, the pH of the solutions was adjusted to lie within the range of 3 to 9. Using a PHM 84 pH metre equipped with glass REF 451 and calomel pHG 201-8 electrodes, this was accomplished. For every pH measurement, a unique experimental setup was established. The suspensions were agitated for 24 h on an agitator operating at 250 revolutions per minute and a temperature of 25 °C. Each sample was filtered using a paper filter with a pore size of 0.5 micrometres. The pH of the subsequent filtrate was then measured. The concentrations of NH4+, NO3, and PO43− were measured using the previously described method.
Equation (1) was utilised to determine the equilibrium adsorption capacity:
q e = V C 0 C e m
where qe is the adsorption capacity at equilibrium (mg·g−1); V is the volume of the solution (dm3); C0 and Ce are the initial and equilibrium concentrations of NH4+, NO3, and PO43− (mg·dm−3); and m is the weight of the adsorbent (g).

2.3.3. Adsorption Isotherms

In this experiment, we added the BCs, HCs, and AC (0.5 g) to 250 cm3 Erlenmeyer flasks containing NH4+, NO3, and PO43− solutions. The solutions had varying concentrations of 25, 50, 100, 200, and 500 mg dm−3, and the pH was kept constant at 7. The suspensions underwent agitation on a shaker operating at 200 rotations per minute and a temperature of 25 °C for a duration of 24 h. The filtration process involved the use of a paper filter with a pore size of 0.5 micrometres to filter the suspensions. The filtrates were subjected to the measurement of NH4+, NO3, and PO43− concentrations.
The adsorption data of the NH4+, NO3, and PO43− on the BCs, HCs, and AC were analysed using the Langmuir and Freundlich isotherm models. The Langmuir model is described in Equation (2) [25]:
C e q e = 1 K L q m + C e q m
where Ce is the equilibrium concentration (mg·dm−3), and constant qm (mg·g−1) and KL are the characteristics of the Langmuir equation (dm3·mg−1) and can be determined from the linearised form (plots of Ce/qe vs. Ce). The Freundlich model is expressed according to Equation (3) [26]:
l n q e = l n K F + 1 n l n C e
where KF is the Freundlich adsorption capacity (mg·g−1), and 1/n is the Freundlich constant. The above equation can be linearised to calculate the parameters KF and n (plots of log qe vs. log Ce).
Studies of the thermodynamics and the effect of temperature are presented in Section 3.6. The experiments were conducted as follows: the BCs and HCs were studied at 20, 35, and 50 °C using initial NH4+, NO3, and PO43− concentrations of 100 mg·dm−3 at pH 7. Each sample was filtered through a 0.5 μm paper filter, and the concentrations of NH4+, NO3, and PO43− were measured.

3. Results and Discussion

3.1. BC and HC Characteristics

Ultimate and proximate analyses of the BC and HC samples produced under various conditions are presented in Table 1. The results indicate that an increase in residence time led to a corresponding increase in the fixed carbon and ash contents, as well as in the O/C and C/N ratios, for the HCs. Conversely, the volatile matter and H/C ratio decreased with an increase in residence time. The results suggest that the residence times of 30 and 180 min did not significantly affect the production of hydrocarbons, as the values obtained were closely aligned. The observed difference in the C/N ratio can be attributed to the increase in N losses that occurs with longer residence times. The findings obtained are consistent with the outcomes and patterns documented in prior research [27,28,29]. The BET method was applied to determine the pore structure. The BET surface area and pore volume of the BCs produced at a low temperature and the HCs are similar, and the values of both parameters increased significantly as the pyrolysis temperature increased, as confirmed in other studies [30,31]. The surface area of the BC stayed at a stable level after passing 500 °C, as presented in Table 1.
The results indicate that an increase in carbonisation temperature led to an increase in the concentrations of fixed carbon and ash, while the volatile matter and atomic ratios (H/C and O/C) decreased for the BCs. The obtained results are consistent with those previously reported in the scientific literature [16,32] for various carbonisation temperatures. According to the findings of previous research, the pyrolysis process primarily results in the transformation of hemicelluloses and cellulose into gaseous products, whereas lignin is predominantly transformed into char at elevated temperatures. The results of this study suggest that increasing the pyrolysis temperature can increase the degree of carbonisation of BC. According to previous research [16], this finding suggests that the BCs produced by this process may have potential benefits for carbon sequestration. Compared to the BCs and HCs, the AC was found to have the highest concentration of fixed carbon and the lowest concentrations of ash and volatile matter [33].
An FTIR spectral analysis plays a crucial role in the identification of distinct functional groups responsible for ion adsorption [18]. As shown in Figure 1, the FTIR spectra of HC30, HC180, BC300, BC500, BC700, and AC were collected and analysed in the presence and absence of NH4+, NO3, and PO43− adsorption.
The vibrational spectra of the HCs and BCs were compared, and it was observed that the number of vibrational peaks was significantly greater in the HC spectra than in the BC spectra. The vibrational peaks at 2850 cm−1, 1698 cm−1, and 1206 cm−1 were attributed to C-H stretching, aromatic C=O, and C-O stretching vibrations, respectively, as depicted in Figure 1. The techniques of hydrolysis and pyrolysis exhibited distinct differences in terms of chemical structure. The observed peaks in the spectra can be attributed to the presence of carboxyl and ester functional groups, as previously reported [34]. The study in [35] reported comparable transmittance values in the range of 4000 to 650 cm−1 for two samples of HCs. This suggests that there were no significant chemical structural modifications resulting from an increase in the residence period.
A correlation was observed between the pyrolysis temperature and the presence of functional groups in the BCs, with a decrease in the number of functional groups as the temperature increased. The obtained spectra exhibited distinct peaks at specific wavenumbers, namely, 2907 cm−1 for methyl CH stretching vibrations, 1703 cm−1 for aromatic carbonyl/carboxyl C=O, 1595 cm−1 for aromatic C=C and C=O, 1028 cm−1 for aliphatic COC and OH, and 814 cm−1 for aromatic CH [36]. The observed shifts in the appearance of the bands are consistent with the findings of previous studies [37,38] and were attributed to the increase in pyrolysis temperature. According to Zhou and colleagues [18], the peaks observed in the range of 1000 to 600 cm−1 are associated with the undulating vibration of aromatic C-H. With an increase in pyrolysis temperature, a reduction in the polar functional groups (OH and C-O) was observed, as indicated by the smaller peaks in their respective spectra [16]. Upon the adsorption of NH4+, NO3, and PO43−, both the biochars (BCs) and hydrochars (HCs) exhibited comparable peaks. Nevertheless, the peaks observed in the BC spectra were relatively feeble, whereas those in the HC spectra were relatively robust. Consistent with previous studies [16,39,40], our FTIR spectra analysis revealed the absence of a significant band subsequent to phosphate adsorption. Upon analysing the wavelengths of particular peaks depicted in Figure 1, slight deviations were observed, consistent with findings reported previously [41,42]. The present study demonstrates the complexation phenomenon between ammonium ions and ionised -OH groups, as well as the bonded -OH bands of carboxylic acids. This was evidenced by a discernible shift in the wavenumber of multiple peaks within the range of 3500 to 3200 cm−1. The confirmation of the ion-exchange phenomenon between the protons of aliphatic C-H and the symmetric stretching vibration of CH2 with ammonium ions has been established through shifts observed at 2920 and 2860 cm−1, as previously reported [42].

3.2. Zeta Potential

The zeta potential values of the HCs, BCs, and AC were measured after the adsorption of NH4+, NO3, and PO43−, as presented in Figure 2. The samples were analysed for their zeta potential at pH 7, which was found to range from −18.7 to −32.6 mV. This suggests that the particles present in the samples possessed negatively charged surfaces. The results of this study indicate that the AC exhibited the highest average zeta potential, while the HC displayed the lowest. An increase in the pH of the suspension resulted in an elevated discrepancy in the zeta potential between the systems. The findings of this study indicate that HCs, BCs, and AC have the ability to selectively adsorb NH4+, NO3, and PO43−. Moreover, it was observed that the influence of specific adsorption became more pronounced at elevated suspension pH levels. Similar findings [16,43] were reported in studies on the utilisation of crop straw feedstock-based biochar for the adsorption of copper and phosphate, respectively.
Furthermore, the zeta potential of the HCs, BCs, and AC became more negative with increased pH values, suggesting that the amount of negative charge increased with the increase in pH and that the adsorption capacity should be lower. The electrostatic adsorption of ions on solid charged surfaces does not tend to affect the surface charge or surface potential of colloidal particles, as the adsorbed ions exist in the diffuse layer of the electric double layers on the particles. Nevertheless, the specific adsorption of ions changes the surface charge and the surface potential of colloidal particles, since these ions go into the Stern layer of the electric double layers and form chemical bonds with the solid particle surfaces. In addition, during the process of cation specific adsorption, some positive charges are transferred to the surface of biochars. This process could lead to a less negative surface charge or even a net positive charge [44]. Accordingly, the zeta potential of the BCs became less negative or changed from negative to positive.

3.3. Adsorption of Ammonium

The impact of carbonisation temperature and residence duration on the adsorption capacity and surface area of the chars was highly significant. The equilibrium isotherms and the adsorption capacity variation as a function of the char used and the initial NH4+ concentration are illustrated in Figure 3. The results of this study indicate that HC180 exhibited a higher adsorption capacity compared to HC30. However, BC700 demonstrated a lower adsorption capacity in comparison to BC500 and BC300. The results indicate that BC300 exhibited the most significant NH4+ adsorption capacity, while HC30 demonstrated the least. Upon increasing the initial concentration of NH4+ from 25 to 500 mg·dm−3, the observed mean values for the HCs exhibited a range of 0.63 to 2.67 mg·g−1, while the mean values for the BCs ranged from 2.59 to 9.05 mg·g−1. In terms of NH4+ adsorption capacity, the AC was surpassed only by BC300.
The adsorption behaviour of the BCs was observed to be influenced by the initial concentration of NH4+ in the solution, as well as the temperature of carbonisation. The results of the observations indicate an inverse relationship between NH4+ adsorption capacity and carbonisation temperature. Specifically, as the temperature of carbonisation increased, the capacity for NH4+ adsorption decreased. Comparable results were obtained by Wang et al. [14] on biochars derived from oak sawdust through pyrolysis at temperatures between 300 °C and 600 °C. The results of this study indicate that the adsorption of NH4+ is augmented by BCs that are generated at lower temperatures. As per prior research, it can be observed in Figure 1 that the BC produced at a higher pyrolysis temperature lacked the presence of aromatic C=O and C=C, -CH2-, CO, and CC functional groups. As a result, it was observed that the functional polar groups present in the material underwent a reaction with NH4+ during the adsorption process. This led to a significant increase in the adsorption capacity of the BC that was produced at lower temperatures, even its surface area was greatly lower than that produced at higher temperatures.
The observed results indicate that there was no significant variation in the NH4+ adsorption capacity of the HCs that were assessed. The results suggest that varying the residence time between 30 and 180 min does not significantly impact the NH4+ adsorption capacity. In their study, Takaya et al. [11] conducted hydrothermal carbonisation at 250 °C to synthesise HC and compared it with BCs obtained from different biomass sources at low and high pyrolysis temperatures (400–450 °C and 600–650 °C, respectively). The study found that there were variations in the physicochemical properties of the char and the refining conditions. However, the NH4+ sorption capacities ranged between 105.8 and 146.4 mg·g−1. As per the findings of the study, it can be inferred that the NH4+ adsorption capacity is not primarily influenced by the surface area [11,45].
The present study employed the Langmuir and Freundlich isotherm models to investigate the adsorption behaviour of NH4+, NO3, and PO43− on various samples of BCs, HCs, and AC. The results of the analysis are presented in Table 2. According to the existing literature, the Langmuir isotherm model posits that sorption occurs uniformly across all binding sites. Conversely, the Freundlich isotherm model suggests that the surface is heterogeneous, with an uneven distribution of adsorption capacity over the surface [46]. The Langmuir model demonstrates superior descriptive capabilities for the adsorption process of NH4+, NO3, and PO43− on the BCs, HCs, and AC, as evidenced by the higher R2 values. Through the utilisation of both BCs and HCs, it was found in the study in [11] that the Langmuir isotherm model exhibited a higher degree of accuracy in describing NH4+ adsorption in comparison to the Freundlich isotherm model.

3.4. Adsorption of Nitrate

The maximum and the minimum NO3 adsorption values were reached by the AC and HC30, respectively (Figure 4). The average values ranged from 0.11 to 0.39 mg·g−1, and 0.59 to 3.46 mg·g−1, respectively, when increasing the initial NO3 concentration from 25 to 500 mg·dm−3.
This study revealed that the adsorption capacity of NO3 was comparatively higher in BC700 as compared to in BC500 and BC300. This indicates that an increase in pyrolysis temperature leads to an improvement in the NO3 adsorption capacity of BCs. In a previous study [14], a range of BCs were analysed. These BCs were produced through the pyrolysis of oak sawdust at different temperatures, ranging from 300 °C to 600 °C.
The results of this study indicate that higher pyrolysis temperatures led to an improvement in the adsorption of NO3. The results indicate that HC180 exhibited a notably greater adsorption capacity compared to HC30 for the target compounds. This suggests that an increase in residence time from 30 to 180 min led to an increase in the quantity of NO3 that could be adsorbed by the material.

3.5. Adsorption of Phosphate

Similarly to the NO3 adsorption capacity using the studied chars, the maximum and the minimum PO43− adsorption values were reached by the AC and HC30, respectively (Figure 5). The average values ranged from 0.12 to 1.23 mg·g−1, and 0.57 to 7.15 mg·g−1, respectively, when increasing the initial PO43− concentration from 25 to 500 mg·dm−3.
The adsorption capacity results for PO43− in the BCs were found to be comparable to those observed for NH4+ and NO3. It was observed that the adsorption capacity was relatively higher for BC700 as compared to BC500 and BC300. This finding further supports the earlier conclusion that a reduction in BC pyrolysis temperature results in an increase in the PO43− adsorption capacity.
The findings of Trazzi et al. [16] indicate that the adsorption and desorption behaviour of PO43− on BC derived from sugar cane bagasse and Miscanthus x giganteus grass is contingent upon the initial concentration of PO43−, pyrolysis temperature, and residence duration. Likewise, Peng et al. [47] found that the adsorption capacity of PO43− was higher when utilising BC obtained through the pyrolysis of sawdust from Pinus trees at a temperature of 750 °C as compared to that produced at 550 °C.
Regarding the HCs, it was observed that HC180 demonstrated a significantly higher adsorption capacity in comparison to HC30. The findings indicate that the augmentation of the residency period from 30 min to 180 min resulted in a rise in the quantity of adsorbed PO43−. Dai et al. [20] were successful in obtaining higher values compared to our investigation by utilising an engineered HC that was synthesised through the HTC of lanthanum-pre-treated rice straw. In a pH range spanning from 2.5 to 10.5, the adsorption capacities of P on the HC in question were observed to exceed 50 mg·g−1. Furthermore, the impact of competing anions (such as Cl, NO3, and SO42−) on P adsorption was found to be negligible [20].
In order to compare the properties of BC and HC, the study in [11] analysed the sorption capacities of PO43− in BCs and HCs that were prepared at different temperatures using different types of waste materials. The PO43− sorption capacities of these materials ranged from 0 to 30 mg·g−1. The study conducted by the researchers revealed that the alterations in the pyrolysis temperature of the BC and the composition of the feedstock did not have a substantial effect on the capacity of char adsorption. Additionally, the study found that the HC generally exhibited a lower adsorption capacity than the BCs. The augmentation of the initial ion concentration has been observed to intensify the adsorption process, plausibly owing to the amplified concentration gradients, thereby leading to a more comprehensive occupation of the responsive adsorption locales [45]. Previous research has indicated that the adsorption of PO43− increased with an increase in the initial concentration of the ion. However, the efficiency of adsorption decreased, likely due to a reduction in the number of active adsorption sites available at higher initial PO43− concentrations, as demonstrated by other studies [48].
The Langmuir model demonstrated superior R2 values for PO43− adsorption, suggesting that this model could more precisely depict the adsorption on the BCs, HCs, and AC, as shown in Table 2. Comparable outcomes were identified in the existing literature for alternative BC [49]. The findings of Takaya et al. [11] suggest that the linearised Freundlich isotherm model provided a slightly better description of the adsorption mechanism. This conclusion was based on the higher R2 values and stronger correlation observed between the experimental and calculated qe values.

3.6. Adsorption Thermodynamics

In the design of sorption systems, two categories of thermodynamic properties are essential: directly measurable properties, like temperature and the equilibrium constant, and properties that cannot be directly measured, such as entropy (S) and free energy (G). To determine the thermodynamic properties that cannot be directly measured, the well-known thermodynamic relation Gibbs free energy (G) is employed. Both Gibbs free energy and entropy play a crucial role in determining the characteristics of sorption chemical reactions or the overall nature of the sorption process. Evaluating these thermodynamic parameters is instrumental in gaining insights into the nature of the adsorption process [50]. The differential heat of adsorption was calculated using the Clausius–Clapeyron expression in the form
H = 2.303 R T 1 T 2 T 1 T 2 l o g C 2 l o g C 1
where C2 and C1 are the equilibrium concentrations at temperatures T2 and T1, respectively. For each adsorption isotherm, we obtained C2 and C1 at a fixed qe value using the Langmuir isotherm expression. Then, the obtained C2 and C1 values were substituted in Equation (4) to obtain the differential heat of adsorption. Equation (5) can be used to obtain the adsorption isosteric heat, which is a positive quantity and is given by the following relationship:
qst = −ΔH
By employing other thermodynamic parameters, including the Gibbs free energy, ΔG, and the entropy change, ΔS was obtained from the following thermodynamic relation:
TΔS = ΔH − ΔG
The free energy change can be calculated using the expression
ΔG = RT lnCe/Cs
where Cs is the solubility of the solute, and it depends on the adsorption temperature. The free energy change obtained using Equation (7) is independent of the temperature within the range of temperatures studied.
Figure 6a–i show the variation in the free energy change, change in the heat content, and entropy change for the transfer of the solutes (single components) ammonium, nitrate, and phosphate from the bulk solution to the adsorbed state on the adsorbent surface of BC700, HC180, and the AC at 50 °C for all the studied systems. Irrespective of the solute and adsorbent type, we observe a general trend for the variation in the thermodynamic parameters as a function of the amount adsorbed. The Gibbs free energy increases with coverage. This can be expected as the equilibrium concentration increases with the adsorption at a fixed T. The differential heat of adsorption decreases with coverage or adsorption, qe. At lower qe, the adsorption is dominated by solute–solid interactions; however, as the adsorption or the coverage increases, the adsorption is dictated by both solute–solute and solute–solid interactions, and this explains the decrease in the ΔH value with the increase in qe. The solid–solute interactions at lower surface adsorption can be due to the adsorption onto the active sites with the highest binding affinity; these typically arise from pores of a size that allows them to commensurate with solute ions or the functional groups. For the case of entropy, it clearly decreases with the increase in the surface coverage. This is an expected observation as the system moves from a disordered state to a more ordered adsorbed state. From the differential heat of adsorption, we calculate the adsorption isosteric heat at zero coverage by estimating that qst = −ΔH at qe = 0. The calculated qst values at zero coverage for BC700, HC180, and the AC are plotted as a function of the solute. In Figure 7, it can be seen that HC180 has strong adsorption towards nitrate ions. This does not mean that this adsorbent possesses the highest adsorption capacity, but it has the potential to target the solute molecules at lower concentrations. The adsorption isosteric heat ranges from 5.6 to 27.5 kJ/mol, 15 to 53.6 kJ/mol, and 3.5 to 23.8 kJ/mol for the biochar, hydrochar, and activated carbon, respectively. According to qst at zero coverage, BC700 shows a strong affinity towards ammonium, followed by nitrate and phosphate. For the case of nitrate and phosphate removal, HC180 shows a stronger affinity towards this solute, followed by BC700 and then AC. Another noteworthy observation is that the qst of BA700 for ammonium and HC180 for nitrate is around 28 and 54 kJ/mol, respectively. These values indicate that adsorption proceeds via chemisorption during the early stage of adsorption.
A large number of previous studies have indicated that the adsorption of ammonium [51,52,53], nitrate [53,54,55,56], and phosphate [57,58,59] is typically exothermic, agreeing with our results (note that the ΔG values obtained in this work are independent of the temperature).
To gain additional insights into Figure 6, we plotted the adsorption capacity of NH4+, NO3, and PO43−, which exhibit an upward trend with an increase in the process temperature in Figure 7. Based on the data, it can be inferred that the adsorption of all three ions onto BC700, HC180, and the AC was facilitated by higher adsorption temperatures. This clearly shows no correlation with qst at zero coverage. This can be expected, as the adsorption at higher concentrations is independent of the binding affinity of the adsorption at zero coverage. Nevertheless, Figure 7 delivers an important message that higher temperatures promote more favourable adsorption, which agrees with some of the earlier works [60,61].

3.7. Comparison with Other Adsorbents

Table 3 presents the adsorption of NH4+, NO3, and PO43− by various BCs and HCs. Our research findings indicate that the quantity of adsorbed ions was influenced by the parameters of the carbonisation process and the initial ion concentrations present in the solution. When evaluating the sorption capacities of different carbonaceous materials, it is imperative to consider these parameters.
Based on the data in Table 3, it appears that the BCs and HCs employed in this study share similar sorption capacities with other adsorbents used in earlier investigations. Higher NH4+ adsorption was observed in HCs derived from oak [11] and HCs activated with KOH [62] compared to HCs derived from other biomass sources. Our HC30s from sugarcane bagasse adsorbed far less ammonium than those in Table 3. When compared to wheat straw BC [34], BC500 was at least ten times more efficient in removing nitrates from the environment.
The sorption capacities of the BCs and HCs examined here are comparable to those of previously investigated adsorbents, as shown in Table 3. Compared to HCs generated from other biomass sources, HCs activated with KOH and HCs derived from oak exhibited much greater NH4+ adsorption. Table 3 displays the ammonium adsorption capacities of several HCs. Sugarcane bagasse HC30 exhibited the lowest value. BC500 was shown to be at least ten times more effective than wheat straw BC at adsorbing ammonium.
Table 3 shows that the BCs and HCs used here compare favourably to other adsorbents in terms of their sorption capacities. KOH-activated HC and oak HC were shown to have much greater NH4+ adsorption than HCs generated from other biomass sources. Table 3 shows that, compared to the other HCs, sugarcane bagasse HC30 was much less effective at adsorbing ammonium. BC500 was at least 10 times more effective at adsorbing ammonium than BC produced from wheat straw.
For the best NO3 adsorption results, use BC [55] produced from refined sugarcane bagasse. In acidic solutions, adsorption was shown to be more efficient than in basic ones. Table 3 suggests that, among the BCs, BC500 had the highest nitrate adsorption rate.
The highest PO43− adsorption capacity of the modified sugarcane BC was achieved through slow pyrolysis at 350 °C with Al enrichment, similar to what was carried out in a prior work [66]. The BCs’ improved P adsorption capacity can be attributed to the material being doped with Al cations, which facilitate bonding between ions and carboxylic groups. Surface irregularity and specific surface area are proven to have a major impact on BCs’ adsorption performance [67]. According to Table 3, alternative HCs, such as sewage sludge HC, KOH-activated sewage sludge HC, and oak-based HC, are more efficacious than the sugarcane bagasse HC30 used in the present study [66]. As a result, the adsorption capacity differs substantially depending on the type of biomass, the degree to which it was pre-treated or activated, and the conditions under which it was manufactured.

4. Conclusions

The utilisation of BCs and HCs has experienced a significant surge owing to their demonstrated efficacy and versatility in various interdisciplinary domains. In the present study, sugarcane bagasse underwent pyrolysis and hydrothermal carbonisation at various temperatures to yield BC and HC. As the pyrolysis temperature was elevated for the BCs, there was a noticeable increase in the fixed carbon, surface area, pore volume, and ash content, whereas the volatile matter and the H/C and O/C atomic ratios experienced a decrease. The observed trend indicated an increase in the fixed carbon content, ash content, and O/C and C/N ratios, as well as volatile matter and H/C ratios, with an increase in residence time for the hydrocarbons. Upon the adsorption of NH4+, NO3, and PO43−, a reduction in the intensities of carboxyl and ester functional groups was observed in the FTIR spectra of both the BCs and HCs. Furthermore, the research findings demonstrate that the Langmuir isotherm model exhibited a marginally better conformity than the Freundlich isotherm model, implying that the adsorption process was homogeneous across all adsorption sites. At elevated temperatures, the exothermic reaction pertaining to the adsorption of the three ions was expedited. The adsorption capacity of the BC and HC adsorbents was observed to be significantly influenced by the type of biomass employed, the extent of pre-treatment or activation of the chars, and the preparation parameters. However, in the case of ammonia adsorption on the BCs, the functional polar groups present in the material underwent a reaction with NH4+ during the adsorption process and had a much stronger effect than the surface area.

Author Contributions

Conceptualization, P.A.T. and W.K.; Methodology, A.S. and J.J.L.; Formal analysis, A.S.; Investigation, J.N.; Writing—original draft, P.A.T. and M.V.; Writing—review & editing, J.N., V.K.K. and W.K.; Supervision, W.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

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

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Kambo, H.S.; Dutta, A. A comparative review of biochar and hydrochar in terms of production, physico-chemical properties and applications. Renew. Sustain. Energy Rev. 2015, 45, 359–378. [Google Scholar]
  2. Atallah, E.; Kwapinski, W.; Ahmad, M.N.; Leahy, J.J.; Zeaiter, J. Effect of water-sludge ratio and reaction time on the hydrothermal carbonization of olive oil mill wastewater treatment: Hydrochar characterization. J. Water Process Eng. 2019, 31, 100813. [Google Scholar] [CrossRef]
  3. Paulo, A.T.; Higa, A.R.; Jeferson, D.; Salvio, M.A.; Rosana Clara Vitoria, H. Biochar: Reality and potential use in forestry. Cienc. Florest. 2018, 28, 875–887. [Google Scholar]
  4. Lua, A.C.; Yang, T.; Guo, J. Effects of pyrolysis conditions on the properties of activated carbons prepared from pistachio-nut shells. J. Anal. Appl. Pyrolysis 2004, 72, 279–287. [Google Scholar] [CrossRef]
  5. Spokas, K.A.; Cantrell, K.B.; Novak, J.M.; Archer, D.W.; Ippolito, J.A.; Collins, H.P.; Boateng, A.A.; Lima, I.M.; Lamb, M.C.; McAloon, A.J.; et al. Biochar: A Synthesis of its Agronomic Impact beyond Carbon Sequestration. J. Environ. Qual. 2012, 41, 973–989. [Google Scholar] [CrossRef]
  6. Joseph, S.D.; Camps-Arbestain, M.; Lin, Y.; Munroe, P.; Chia, C.; Hook, J.; Van Zwieten, L.; Kimber, S.; Cowie, A.; Singh, B.; et al. An investigation into the reactions of biochar in soil. Aust. J. Soil Res. 2010, 48, 501–515. [Google Scholar] [CrossRef]
  7. Morales, M.M.; Comerford, N.; Guerrini, I.A.; Falcão, N.P.S.; Reeves, J.B. Sorption and desorption of phosphate on biochar and biochar-soil mixtures. Soil Use Manag. 2013, 29, 306–314. [Google Scholar] [CrossRef]
  8. Mumme, J.; Eckervogt, L.; Pielert, J.; Diakité, M.; Rupp, F.; Kern, J. Hydrothermal carbonization of anaerobically digested maize silage. Bioresour. Technol. 2011, 102, 9255–9260. [Google Scholar] [CrossRef]
  9. Hoekman, S.K.; Broch, A.; Robbins, C. Hydrothermal carbonization (HTC) of lignocellulosic biomass. Energy Fuels 2011, 25, 1802–1810. [Google Scholar] [CrossRef]
  10. Zhang, Z.; Zhu, Z.; Shen, B.; Liu, L. Insights into biochar and hydrochar production and applications: A review. Energy 2019, 171, 581–598. [Google Scholar] [CrossRef]
  11. Takaya, C.A.; Fletcher, L.A.; Singh, S.; Anyikude, K.U.; Ross, A.B. Phosphate and ammonium sorption capacity of biochar and hydrochar from different wastes. Chemosphere 2016, 145, 518–527. [Google Scholar] [CrossRef]
  12. Mengistu, D.; Nilsen, V.; Heistad, A.; Kvaal, K. Detection and quantification of tire particles in sediments using a combination of simultaneous thermal analysis, fourier transform infra-red, and parallel factor analysis. Int. J. Environ. Res. Public Health 2019, 16, 3444. [Google Scholar] [CrossRef]
  13. Yao, Y.; Gao, B.; Zhang, M.; Inyang, M.; Zimmerman, A.R. Effect of biochar amendment on sorption and leaching of nitrate, ammonium, and phosphate in a sandy soil. Chemosphere 2012, 89, 1467–1471. [Google Scholar] [CrossRef]
  14. 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]
  15. Hale, S.E.; Alling, V.; Martinsen, V.; Mulder, J.; Breedveld, G.D.; Cornelissen, G. The sorption and desorption of phosphate-P, ammonium-N and nitrate-N in cacao shell and corn cob biochars. Chemosphere 2013, 91, 1612–1619. [Google Scholar] [CrossRef]
  16. Trazzi, P.A.; Leahy, J.J.; Hayes, M.H.B.; Kwapinski, W. Adsorption and desorption of phosphate on biochars. J. Environ. Chem. Eng. 2016, 4, 37–46. [Google Scholar] [CrossRef]
  17. Kameyama, K.; Miyamoto, T.; Shiono, T.; Shinogi, Y. Influence of Sugarcane Bagasse-derived Biochar Application on Nitrate Leaching in Calcaric Dark Red Soil. J. Environ. Qual. 2012, 41, 1131–1137. [Google Scholar] [CrossRef]
  18. Zhou, Z.H.; Yuan, J.; Hu, M. Adsorption of ammonium from aqueous solutions on environmentally friendly barbecue bamboo charcoal: Characteristics and kinetic and thermodynamic studies. Environ. Prog. Sustain. Energy 2015, 34, 655–662. [Google Scholar] [CrossRef]
  19. Parvage, M.M.; Ulén, B.; Eriksson, J.; Strock, J.; Kirchmann, H. Phosphorus availability in soils amended with wheat residue char. Biol. Fertil. Soils 2013, 49, 245–250. [Google Scholar] [CrossRef]
  20. Dai, L.; Wu, B.; Tan, F.; He, M.; Wang, W.; Qin, H.; Tang, X.; Zhu, Q.; Pan, K.; Hu, Q. Engineered hydrochar composites for phosphorus removal/recovery: Lanthanum doped hydrochar prepared by hydrothermal carbonization of lanthanum pretreated rice straw. Bioresour. Technol. 2014, 161, 327–332. [Google Scholar] [CrossRef]
  21. ICS 75.160.10, DD CEN/TS 14774-3:2009; Solid biofuels—Methods for the Determination of Moisture Content—Oven Dry Method—Part 3: Moisture in General Analysis Sample. Comite Europeen de Normalisation: Brussels, Belgium, 2009.
  22. ICS 75.160.10, DD CEN/TS 14775:2009; Solid Biofuels—Determination of Ash Content. Comite Europeen de Normalisation: Brussels, Belgium, 2009.
  23. ICS 75.160.10, DD CEN/TS 15148:2009; Solid Biofuels—Determination of the Content of Volatile Matter. Comite Europeen de Normalisation: Brussels, Belgium, 2009.
  24. Langmuir, I. The constitution and fundamental properties of solids and liquids. Part I. Solids. J. Am. Chem. Soc. 1916, 38, 2221–2295. [Google Scholar] [CrossRef]
  25. Freundlich, H. Über die Adsorption in Lösungen. Z. für Phys. Chem. 1907, 57U, 385–470. [Google Scholar] [CrossRef]
  26. Kołodyńska, D.; Wnętrzak, R.; Leahy, J.J.; Hayes MH, B.; Kwapiński, W.; Hubicki, Z. Kinetic and adsorptive characterization of biochar in metal ions removal. Chem. Eng. J. 2012, 197, 295–305. [Google Scholar] [CrossRef]
  27. Fang, J.; Gao, B.; Chen, J.; Zimmerman, A.R. Hydrochars derived from plant biomass under various conditions: Characterization and potential applications and impacts. Chem. Eng. J. 2015, 267, 253–259. [Google Scholar] [CrossRef]
  28. Poerschmann, J.; Weiner, B.; Wedwitschka, H.; Zehnsdorf, A.; Koehler, R.; Kopinke, F.-D. Characterization of biochars and dissolved organic matter phases obtained upon hydrothermal carbonization of Elodea nuttallii. Bioresour. Technol. 2015, 189, 145–153. [Google Scholar] [CrossRef]
  29. Reza, M.T.; Rottler, E.; Tölle, R.; Werner, M.; Ramm, P.; Mumme, J. Production, characterization, and biogas application of magnetic hydrochar from cellulose. Bioresour. Technol. 2015, 186, 34–43. [Google Scholar] [CrossRef]
  30. Jamilatun, S.; Amelia, S.; Pitoyo, J.; Ma’Arif, A.; Mufandi, I. Preparation and characteristics of effective biochar derived from sugarcane bagasse as adsorbent. Int. J. Renew. Ener. Res. 2023, 13, 673–680. [Google Scholar]
  31. Melligan, F.; Dussan, K.; Auccaise, R.; Novotny, E.H.; Leahy, J.J.; Hayes, M.H.B.; Kwapinski, W. Characterisation of the products from pyrolysis of residues after acid hydrolysis of Miscanthus. Bioresour. Technol. 2012, 108, 258–263. [Google Scholar] [CrossRef]
  32. Song, W.; Guo, M. Quality variations of poultry litter biochar generated at different pyrolysis temperatures. J. Anal. Appl. Pyrolysis 2012, 94, 138–145. [Google Scholar] [CrossRef]
  33. Brown, R. Biochar for Environmental Management. In Biochar for Environmental Management; Lehmann, J., Joseph, S., Eds.; Earthscan: London, UK, 2012. [Google Scholar]
  34. Fuertes, A.B.; Arbestain, M.C.; Sevilla, M.; Maciá-Agulló, J.A.; Fiol, S.; López, R.; Smernik, R.J.; Aitkenhead, W.P.; Arce, F.; Macias, F. Chemical and structural properties of carbonaceous products obtained by pyrolysis and hydrothermal carbonisation of corn stover. Aust. J. Soil Res. 2010, 48, 618–626. [Google Scholar] [CrossRef]
  35. Zhang, Q. Production of Fuel Ethanol from Hydrotherm-Pretreated Corn Stover by Pichia stipitis. Adv. Mater. Res. 2013, 641–642, 943–946. [Google Scholar] [CrossRef]
  36. Grams, J. Surface analysis of solid products of thermal treatment of lignocellulosic biomass. J. Anal. Sci. 2022, 161, 105429. [Google Scholar] [CrossRef]
  37. Gai, X.; Wang, H.; Liu, J.; Zhai, L.; Liu, S.; Ren, T.; Liu, H. Effects of Feedstock and Pyrolysis Temperature on Biochar Adsorption of Ammonium and Nitrate. PLoS ONE 2014, 9, e113888. [Google Scholar] [CrossRef]
  38. Chen, B.; Zhou, D.; Zhu, L. Transitional adsorption and partition of nonpolar and polar aromatic contaminants by biochars of pine needles with different pyrolytic temperatures. Environ. Sci. Technol. 2008, 42, 5137–5143. [Google Scholar] [CrossRef]
  39. Zeng, Z.; Zhang, S.D.; Li, T.G.; Zhao, F.L.; He, Z.L.; Zhao, H.P.; Yang, X.E.; Wang, H.L.; Zhao, J.; Rafiq, M.T. Sorption of ammonium and phosphate from aqueous solution by biochar derived from phytoremediation plants. J. Zhejiang Univ. Sci. B 2013, 14, 1152. [Google Scholar] [CrossRef]
  40. Halajnia, A.; Oustan, S.; Najafi, N.; Khataee, A.R.; Lakzian, A. Adsorption–desorption characteristics of nitrate, phosphate and sulfate on Mg–Al layered double hydroxide. Appl. Clay Sci. 2013, 80–81, 305–312. [Google Scholar] [CrossRef]
  41. Liu, H.; Dong, Y.; Wang, H.; Liu, Y. Ammonium adsorption from aqueous solutions by strawberry leaf powder: Equilibrium, kinetics and effects of coexisting ions. Desalination 2010, 263, 70–75. [Google Scholar] [CrossRef]
  42. Wahab, M.A.; Boubakri, H.; Jellali, S.; Jedidi, N. Characterization of ammonium retention processes onto cactus leaves fibers using FTIR, EDX and SEM analysis. J. Hazard. Mater. 2012, 241–242, 101–109. [Google Scholar] [CrossRef]
  43. Tong, X.J.; Li, J.Y.; Yuan, J.H.; Xu, R.K. Adsorption of Cu(II) by biochars generated from three crop straws. Chem. Eng. J. 2011, 172, 828–834. [Google Scholar] [CrossRef]
  44. Xu, R.K.; Xiao, S.C.; Yuan, J.H.; Zhao, A.Z. Adsorption of methyl violet from aqueous solutions by the biochars derived from crop residues. Bioresour. Technol. 2011, 102, 10293–10298. [Google Scholar] [CrossRef]
  45. Zakutevskyy, O.; Shaposhnikova, T.; Kotynska, L.; Biedrzycka, A.; Skwarek, E. Sorption of ammonium and phosphate ions from aqueous solutions by carbon and mineral sorbents. Physicochem. Probl. Miner. Process. 2022, 58, 150285. [Google Scholar] [CrossRef]
  46. Wasewar, K.L.; Atif, M.; Prasad, B.; Mishra, I.M. Adsorption of zinc using tea factory waste: Kinetics, equilibrium and thermodynamics. Clean Soil Air Water 2008, 36, 320–329. [Google Scholar] [CrossRef]
  47. Peng, F.; He, P.W.; Luo, Y.; Lu, X.; Liang, Y.; Fu, J. Adsorption of Phosphate by Biomass Char Deriving from Fast Pyrolysis of Biomass Waste. Clean Soil Air Water 2012, 40, 493–498. [Google Scholar] [CrossRef]
  48. Chen, B.; Chen, Z.; Lv, S. A novel magnetic biochar efficiently sorbs organic pollutants and phosphate. Bioresour. Technol. 2011, 102, 716–723. [Google Scholar] [CrossRef]
  49. Kilpimaa, S.; Runtti, H.; Kangas, T.; Lassi, U.; Kuokkanen, T. Removal of phosphate and nitrate over a modified carbon residue from biomass gasification. Chem. Eng. Res. Des. 2014, 92, 1923–1933. [Google Scholar] [CrossRef]
  50. Kumar, K.V.; Gadipelli, S.; Wood, B.; Ramisetty, K.A.; Stewart, A.A.; Howard, C.A.; Brett, D.J.L.; Rodriguez-Reinoso, F. Characterization of the adsorption site energies and heterogeneous surfaces of porous materials. J. Mater. Chem. A 2019, 17, 10104–10137. [Google Scholar] [CrossRef]
  51. Singh, K.; Dixit, U.; Mohan, S. Comparative study of adsorption, kinetics, and thermodynamics of selected cationic and anionic surfactants on ultrasound-assisted bagasse. Water Air Soil Pollut. 2024, 235, 10. [Google Scholar] [CrossRef]
  52. Vu, T.M.; Nguyen, T.M.P.; Van, H.T.; Hoang, V.H.; Nga, L.T.Q.; Hoang, L.P.; Ravindran, B. High removal efficiency of ammonium from aqueous solution by colloidal silver nanoparticles: Batch adsorption. Urban Water J. 2023. [Google Scholar] [CrossRef]
  53. Wang, Y.; Song, X.; Xu, Z.; Cao, X.; Song, J.; Huang, W.; Ge, X.; Wang, H. Adsorption of nitrate and ammonium from water simultaneously using composite adsorbents constructed with functionalized biochar and modified zeolite. Water Air Soil Pollut. 2021, 232, 198. [Google Scholar] [CrossRef]
  54. Yao, J.; Wang, Z.; Liu, M.; Bai, B.; Zhang, C. Nitrate-nitrogen adsorption characteristics and mechanisms of various garden waste biochars. Materials 2023, 16, 5726. [Google Scholar] [CrossRef]
  55. Allahkarami, E.; Azadmehr, A.; Noroozi, F.; Farrokhi, S.; Sillanpää, M. Nitrate adsorption onto surface-modified red mud in batch and fixed-bed column systems: Equilibrium, kinetic, and thermodynamic studies. Environ. Sci. Pollut. Res. 2022, 29, 48438–48452. [Google Scholar] [CrossRef]
  56. Bahrami, M.; Amiri, M.J. Nitrate removal from contaminated waters using modified rice husk ash by Hexadecyltrimethylammonium bromide surfactant. React. Kinet. Mech. Catal. 2022, 135, 459–478. [Google Scholar]
  57. Saleh, M.; Alterkaoui, A.; Ozdemir, N.C.; Arslan, H.; Bilici, Z.; Dizge, N. Adsorption of phosphate ions and reactive red 180 from aqueous solution using thermally activated lemon peels waste. Int. J. Environ. Sci. Technol. 2024, 21, 1683–1696. [Google Scholar] [CrossRef]
  58. Eskikaya, O.; Arslan, H.; Gun, M.; Bouchareb, R.; Dizge, N. Adsorption of Direct Orange 46 and phosphate ions on waste tomato stem ash used as a bio-based adsorbent. Environ. Prog. Sustain. Energy 2023, 42, e14192. [Google Scholar] [CrossRef]
  59. Sun, Y.; Gu, Y.; Xiao, S. Adsorption behaviors and mechanisms of Al-Fe dual-decorated biochar adsorbent for phosphate removal from rural wastewater. J. Dispers. Sci. Technol. 2023, 44, 2520–2531. [Google Scholar] [CrossRef]
  60. Huo, H.; Lin, H.; Dong, Y.; Cheng, H.; Wang, H.; Cao, L. Ammonia-nitrogen and phosphates sorption from simulated reclaimed waters by modified clinoptilolite. J. Hazard. Mater. 2012, 229–230, 292–297. [Google Scholar]
  61. Chitrakar, R.; Tezuka, S.; Sonoda, A.; Sakane, K.; Ooi, K.; Hirotsu, T. Adsorption of phosphate from seawater on calcined MgMn-layered double hydroxides. J. Colloid Interface Sci. 2005, 290, 45–51. [Google Scholar] [CrossRef]
  62. Takaya, C.A.; Parmar, K.R.; Fletcher, L.A.; Ross, A.B. Biomass-Derived Carbonaceous Adsorbents for Trapping Ammonia. Agriculture 2019, 9, 16. [Google Scholar] [CrossRef]
  63. Gao, F.; Xue, Y.; Deng, P.; Cheng, X.; Yang, K. Removal of aqueous ammonium by biochars derived from agricultural residuals at different pyrolysis temperatures. Chem. Speciat. Bioavailab. 2015, 27, 92–97. [Google Scholar] [CrossRef]
  64. Zhang, T.; Wu, X.; Fan, X.; Tsang, D.C.W.; Li, G.; Shen, Y. Corn waste valorization to generate activated hydrochar to recover ammonium nitrogen from compost leachate by hydrothermal assisted pretreatment. J. Environ. Manag. 2019, 236, 108–117. [Google Scholar] [CrossRef] [PubMed]
  65. Bakly, S.; Al-Juboori, R.A.; Bowtell, L. Macadamia Nutshell Biochar for Nitrate Removal: Effect of Biochar Preparation and Process Parameters. C 2019, 5, 47. [Google Scholar] [CrossRef]
  66. Kameyama, K.; Miyamoto, T.; Iwata, Y.; Shiono, T. Influences of feedstock and pyrolysis temperature on the nitrate adsorption of biochar. Soil Sci. Plant Nutr. 2016, 62, 180–184. [Google Scholar] [CrossRef]
  67. Yang, J.; Li, H.; Zhang, D.; Wu, M.; Pan, B. Limited role of biochars in nitrogen fixation through nitrate adsorption. Sci. Total Environ. 2017, 592, 758–765. [Google Scholar] [CrossRef]
  68. Hafshejani, L.D.; 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. C 2016, 95, 101–111. [Google Scholar] [CrossRef]
  69. Xue, Y.; Hou, H.; Zhu, S. Characteristics and mechanisms of phosphate adsorption onto basic oxygen furnace slag. J. Hazard. Mater. 2009, 162, 973–980. [Google Scholar] [CrossRef] [PubMed]
  70. Spataru, A.; Spataru, A.; Jain, R.; Chung, J.W.; Gerner, G.; Krebsc, R.; Lens PN, L. Enhanced adsorption of orthophosphate and copper onto hydrochar derived from sewage sludge by KOH activation. RSC Adv. 2016, 6, 101827–101834. [Google Scholar] [CrossRef]
  71. Novais, S.V.; Zenero, M.D.O.; Barreto MS, C.; Montes, C.R.; Cerri, C.E.P. Phosphorus removal from eutrophic water using modified biochar. Sci. Total Environ. 2018, 633, 825–835. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The FTIR spectra of (a) HC30, (b) HC180, (c) BC300, (d) BC500, (e) BC700, and (f) AC before and after NH4+, NO3, and PO43− adsorption.
Figure 1. The FTIR spectra of (a) HC30, (b) HC180, (c) BC300, (d) BC500, (e) BC700, and (f) AC before and after NH4+, NO3, and PO43− adsorption.
Applsci 14 02280 g001
Figure 2. Zeta potential of HCs, BCs, and AC after NH4+, NO3, and PO43− adsorption.
Figure 2. Zeta potential of HCs, BCs, and AC after NH4+, NO3, and PO43− adsorption.
Applsci 14 02280 g002
Figure 3. Equilibrium isotherm plots at 20 °C for NH4+ sorption on BCs, HCs, and AC. Solid and dashed lines represent the Langmuir and Freundlich isotherm data models, respectively.
Figure 3. Equilibrium isotherm plots at 20 °C for NH4+ sorption on BCs, HCs, and AC. Solid and dashed lines represent the Langmuir and Freundlich isotherm data models, respectively.
Applsci 14 02280 g003
Figure 4. Equilibrium isotherm plots at 20 °C for NO3 sorption on BCs, HCs, and AC. Solid and dashed lines represent the Langmuir and Freundlich isotherm data models, respectively.
Figure 4. Equilibrium isotherm plots at 20 °C for NO3 sorption on BCs, HCs, and AC. Solid and dashed lines represent the Langmuir and Freundlich isotherm data models, respectively.
Applsci 14 02280 g004
Figure 5. Equilibrium isotherm plots at 20 °C for PO43− sorption on BCs, HCs, and AC. Solid and dashed lines represent the Langmuir and Freundlich isotherm data models, respectively.
Figure 5. Equilibrium isotherm plots at 20 °C for PO43− sorption on BCs, HCs, and AC. Solid and dashed lines represent the Langmuir and Freundlich isotherm data models, respectively.
Applsci 14 02280 g005
Figure 6. Plot of thermodynamic parameters, ΔG, ΔH, and ΔS for the (a) adsorption of ammonium onto BC700, (b) adsorption of ammonium onto HC180, (c) adsorption of ammonium onto AC, (d) adsorption of nitrate onto BC700, (e) adsorption of nitrate onto HC180, (f) adsorption of nitrate onto AC, (g) adsorption of phosphate onto BC700, (h) adsorption of phosphate onto HC180, (i) adsorption of phosphate onto AC.
Figure 6. Plot of thermodynamic parameters, ΔG, ΔH, and ΔS for the (a) adsorption of ammonium onto BC700, (b) adsorption of ammonium onto HC180, (c) adsorption of ammonium onto AC, (d) adsorption of nitrate onto BC700, (e) adsorption of nitrate onto HC180, (f) adsorption of nitrate onto AC, (g) adsorption of phosphate onto BC700, (h) adsorption of phosphate onto HC180, (i) adsorption of phosphate onto AC.
Applsci 14 02280 g006
Figure 7. Effect of temperature on the sorption of 25, 100, and 500 mg·dm−3 initial concentrations of ammonium (ac), nitrate (df), and phosphate (gi) by BC 700 (a,d,g), HC 180 (b,e,h), and AC (c,f,i).
Figure 7. Effect of temperature on the sorption of 25, 100, and 500 mg·dm−3 initial concentrations of ammonium (ac), nitrate (df), and phosphate (gi) by BC 700 (a,d,g), HC 180 (b,e,h), and AC (c,f,i).
Applsci 14 02280 g007
Table 1. Ultimate and proximate analyses of BC and HC samples produced under different pyrolysis and hydrothermal carbonisation conditions.
Table 1. Ultimate and proximate analyses of BC and HC samples produced under different pyrolysis and hydrothermal carbonisation conditions.
Ultimate Analysis
wt.%
Atomic RatiosBET Surface Area
m2/g
Pore Volume
cm3/g
Proximate Analysis
wt.%
MaterialNCHSOC/NH/CO/C AshVolatile
Matter
Fixed
Carbon
HC300.4967.26.150.019.91380.0920.2971.40.0044.8348.846.4
HC1800.3665.25.770.021.21790.0880.3251.80.0046.3045.947.8
BC3000.3956.45.670.028.31450.1000.5022.40.0036.3058.735.0
BC5000.5769.12.960.013.71210.0430.19980.40.03210.115.874.2
BC7000.5476.31.370.00.971420.0180.01384.30.03017.85.7276.5
AC0.4886.10.300.06.511790.0030.0764001.3653.604.4492.0
Table 2. Langmuir and Freundlich isotherm parameters for the adsorption of ammonium, nitrate, and phosphate on BCs, HCs, and AC.
Table 2. Langmuir and Freundlich isotherm parameters for the adsorption of ammonium, nitrate, and phosphate on BCs, HCs, and AC.
System Material
BC300BC500BC700HC30HC180AC
Ammonium
FreundlichKF1.1771.0160.7550.1920.1861.117
n2.7372.7002.5022.1872.1342.778
R20.8860.8460.8960.8660.9440.886
Langmuirq09.8338.1838.1372.9703.4349.050
KL0.0280.0320.0200.0160.0120.029
R20.9680.9460.9800.9750.9930.967
Nitrate
FreundlichKF0.0360.0730.0910.0330.0930.093
n1.5651.5831.6321.9711.5231.539
R20.6880.6740.6510.8260.7480.708
Langmuirq01.6373.0513.2610.6981.3994.529
KL0.0080.0090.0100.0130.0070.008
R20.8790.8800.8710.9530.9020.893
Phosphate
FreundlichKF0.1190.2720.3010.0330.0980.477
n1.9922.1472.1931.9712.1072.092
R20.8290.8090.8600.8260.8050.832
Langmuirq02.4134.1844.6990.6981.6158.292
KL0.0130.0170.0140.0130.0160.014
R20.9550.9480.9560.9530.9470.943
Table 3. Ammonium, nitrate, and phosphate adsorption capacities of different BCs and HCs formed under different conditions.
Table 3. Ammonium, nitrate, and phosphate adsorption capacities of different BCs and HCs formed under different conditions.
Carbonaceous
Materials
Conditions for the
Adsorbent
Adsorbed IonSolution Concentration
mg dm−3
Adsorption
Capacity
mg g−1
Ref.
Oak-based HCHydrothermal carbonisation 250 °CNH4+436.0[62]
Oak-based BCSlow pyrolysis at 450 °C 9.2
Slow pyrolysis at 650 °C 8.9
Peanut shell, corncob, or
cotton stalk BC
Pyrolysis at 300, 450, or 600 °C for 2 hNH4+5016–18[63]
Wheat straw BCPyrolysis at 500 °C for 1.5 hNH4+500.33[37]
Corn straw BC 0.92
Peanut shell BC 0.54
Sugarcane bagasse BCPyrolysis at 500 °C for 1 hNH4+504.46This study
Sugarcane bagasse HC30Hydrothermal carbonisation at 250 °C for 30 min 1.37
Corn cob HC230 or 260 °C for 30 minNH4+20–3000≈34–107[64]
KOH-activated HCCorn cob HC mixed with 200 mL of 3 M KOH ≈48–140
Oak-based hydrocharHydrothermal carbonisation 250 °CNH4+1000109.7[11]
Oak-based BCSlow pyrolysis at 450 °C 129.4
Slow pyrolysis at 650 °C 123.5
Macadamia nutshell BCPyrolysis at 1000 °C for 1 hNO310 [65]
Sugarcane bagasse BCPyrolysis at 800 °C for 2 hNO3100.06[66]
Bamboo chip BC 0.15
Japanese cedar BC 0.23
Rice husk BC 0.07
Poultry manure BC 0.02
Pine sawdust BCPyrolysis at 500 °C for 4 hNO3300.45[67]
Corn straw BC 0.15
Peanut hull BC 0.38
Rice straw BC 0.15
Wheat straw BCPyrolysis at 500 °C for 1.5 hNO3500.021[37]
Corn straw BC 0.032
Peanut shell BC 0.024
Sugarcane bagasseFeedstockNO3502.1[68]
Sugarcane bagasse BCPyrolysis at 200 to 600 °C for 4 h 11. 6
Modified sugarcane bagasse BCSugarcane bagasse BC added in epichlorohydrin and N-dimethylformamide 28.2
Sugarcane bagasse BCPyrolysis at 500 °C for 1 hNO3501.34This study
Sugarcane bagasse HC30Hydrothermal carbonisation at 250 °C for 30 min 0.35
Pine sawdust BCPyrolysis at 550 °C for 15 minPO43−3010[47]
Pyrolysis at 750 °C for 15 min 14
Activated carbon residue BCDowndraft gasifier 1000 °CPO43−5021[69]
Oak sawdust BCPyrolysis at 500 °C for 30 minPO43−6111[19]
Sugarcane BCPyrolysis at 300 °C for 20 minPO43−1001.63[16]
Pyrolysis at 700 °C for 20 min 10.0
Miscanthus BCPyrolysis at 300 °C for 20 min 1.83
Pyrolysis at 700 °C for 20 min 11.4
Corn BCPyrolysis at 300 °C for 3 hPO43−100≈35[17]
Pyrolysis at 600 °C for 3 h ≈40
Sugarcane bagasse BCPyrolysis at 500 °C for 1 hPO43−1002.76This study
Sugarcane bagasse HC30Hydrothermal carbonisation at 250 °C for 30 min 0.45
Sewage sludge HCHydrothermally carbonised for 5 h at 210 °C 21 and 24 barPO43−1502.3[70]
KOH-activated HCProduced sewage sludge HC was washed with 1 M KOH 14.2
Oak-based hydrocharHydrothermal carbonisation at 250 °CPO43−40026.6[11]
Oak-based BCSlow pyrolysis at 450 °C 5.5
Slow pyrolysis at 650 °C 3.6
Poultry manure BCslowly pyrolysed at 350 °C after Al dopingPO43−3000701[71]
Sugarcane BCslowly pyrolysed at 350 °C after Al doping 759
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

Trazzi, P.A.; Vashishtha, M.; Najser, J.; Schmalenberger, A.; Kannuchamy, V.K.; Leahy, J.J.; Kwapinski, W. Adsorption of Ammonium, Nitrate, and Phosphate on Hydrochars and Biochars. Appl. Sci. 2024, 14, 2280. https://doi.org/10.3390/app14062280

AMA Style

Trazzi PA, Vashishtha M, Najser J, Schmalenberger A, Kannuchamy VK, Leahy JJ, Kwapinski W. Adsorption of Ammonium, Nitrate, and Phosphate on Hydrochars and Biochars. Applied Sciences. 2024; 14(6):2280. https://doi.org/10.3390/app14062280

Chicago/Turabian Style

Trazzi, Paulo André, Mayank Vashishtha, Jan Najser, Achim Schmalenberger, Vasanth Kumar Kannuchamy, James J. Leahy, and Witold Kwapinski. 2024. "Adsorption of Ammonium, Nitrate, and Phosphate on Hydrochars and Biochars" Applied Sciences 14, no. 6: 2280. https://doi.org/10.3390/app14062280

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