Next Article in Journal
Reverse Logistics Network Model of Dual-Channel Recycling Boxes Based on Genetic Algorithm Optimization: A Multi-Objective and Uncertain Environment Perspective
Previous Article in Journal
Community Tourism Conditions and Sustainable Management of a Community Tourism Association: The Case of Cruz Pata, Peru
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Tetracycline Removal from Water by Adsorption on Hydrochar and Hydrochar-Derived Activated Carbon: Performance, Mechanism, and Cost Calculation

1
Faculty of Environment, Thai Nguyen University of Agriculture and Forestry (TUAF), Thai Nguyen City 24000, Vietnam
2
Department of Resource Management, Thai Nguyen University of Agriculture and Forestry (TUAF), Thai Nguyen City 24000, Vietnam
3
Department of Environmental Engineering, Chung Yuan Christian University, Taoyuan 320314, Taiwan
4
Center for Environmental Risk Management, Chung Yuan Christian University, Taoyuan 320314, Taiwan
*
Authors to whom correspondence should be addressed.
Sustainability 2023, 15(5), 4412; https://doi.org/10.3390/su15054412
Submission received: 28 November 2022 / Revised: 15 January 2023 / Accepted: 20 January 2023 / Published: 1 March 2023
(This article belongs to the Section Sustainable Water Management)

Abstract

:
The objective of this study was to investigate the adsorption performance and mechanisms of tetracycline (TC) on hydrochar and hydrochar-derived activated carbon. We also assessed the influence of the solution pH and ionic strength on the adsorption of these compounds and studied their removal by synthetic adsorbents. The maximum adsorption capacities of TC estimated by the Langmuir model in pH 5.5 solution at 25 °C were found to follow the order: ACZ1175 (257.28 mg/g) > MGH (207.11 mg/g) > WAC (197.52 mg/g) > MOPH (168.50 mg/g) > OPH (85.79 mg/g) > GH (75.47 mg/g). The pH value and ionic strength affected TC’s adsorption on the adsorbents. These results indicate that the electrostatic interaction plays a critical role in these adsorption processes. Moreover, adsorption kinetic curves and adsorption isotherms demonstrated that electrostatic interactions were not the only adsorption driving force. Except for electrostatic interactions, the main adsorption mechanisms involved hydrogen bonding and π-π interaction. In addition, the cost of oxidized hydrochar (USD 4.71/kg) is slightly higher than that of hydrochar-derived activated carbon (USD 3.47/kg). This production cost would be lower when it can be produced on a large scale. The outcomes of this study show that the modified-hydrochar and hydrochar-derived activated carbon had the potential for TC removal in wastewater.

1. Introduction

During the 21st century, antibiotics in the environment have received increasing attention, because they might cause antibiotic resistance in microorganisms [1]. Antibiotics are considered to be persistent micro-pollutants in the water environment; most of them cannot be absorbed and metabolized completely by the bodies of humans and animals, which causes a large amount of them to be released into the environment through urine and feces [1,2].
Tetracycline hydrochloride (TC) is one of the most commonly used antibiotics and is also classified as an emerging contaminant. It could be found in soils, surface water, and groundwater, even in drinking water [2]. In general, TC is used in human infection medicine, veterinary medicine, husbandry, and aquaculture growth promoters [3]. Moreover, it might be transferred to the environment via human activities and animal wastes [4]. When TC is widespread in the environment, it can cause a diversity of harmful influences such as chronic toxicity, reduced human immunity, and diffusion in antibiotic-resistant genes [1,5]. Hence, the removal of TC is of great interest and attention in recent years.
A number of methods have been developed to remove TC from an aqueous solution. These technologies include oxidative degradation [6,7], photo-electro catalytic degradation [8,9], biodegradation [10], ozonation [11], and membrane filter [12]. Although these technologies could effectively remove TCs, the technologies frequently need a relatively higher cost and complex operation process.
Adsorption technique is a practical method for the removal of TC from an aqueous solution [13,14,15]. Moreover, adsorption is regarded as a more efficient and cost-effective method. Many investigators have developed new adsorbents for the adsorption of TC, such as graphene oxide [16], aluminum oxide [17], Fe-Mn binary oxide [18], montmorillonite [19], and carbonaceous materials [20,21]. Among these adsorbents, activated carbon (AC) and biochar (BC) have received high interest because they showed the benefit in adsorption amount, cost effectiveness, and operation [5,22,23,24].
A number of investigators have presented the potential adsorption mechanisms between TC and carbonaceous materials including metal bridging, H-bonding, π-π interaction, and van der Waals force [25,26,27]. In a previous work, the electrostatic force was regarded as the primary mechanism controlling the adsorption of organic pollutants (methylene blue) onto the hydrochars [28] and hydrochar-derived activated carbon [29], whereas hydrogen bonding and other mechanisms played minor roles. Hydrochar (HC) has been developed for adsorption in recent years. Only a few studies regarding hydrochar and hydrochar-derived AC to remove TC in wastewater have been presented. The adsorption mechanisms have not been widely discussed in the literature. Thus, the adsorption amount and adsorption mechanisms of TC on the given hydrochar and hydrochar-derived AC need to be further evaluated.
With this background, it is clear that the adsorption of antibiotics on carbon materials has yet to be fully elucidated. Hence, in this study, orange peel and teak sawdust were used as the feedstock of HCs and ACs due to their abundance in agricultural wastes. To improve the adsorption capacities of adsorbents, HNO3 and ZnCl2 were used to activate HCs and enhance the specific surface area. The purposes of this study were (i) to examine the adsorption efficiency and mechanisms of hydrochar and ACs for TC removal; and (ii) to confirm electrostatic interaction between adsorbent and TC. Moreover, the effect of pH on the adsorption capacity of TC was studied in detail and adsorption mechanisms of TC on each adsorbent were elucidated. Finally, this research can address the key relevant challenges for the future applications of hydrochar and hydrochar-derived AC.

2. Materials and Methods

2.1. Tetracycline Characterization

In this study, tetracycline hydrochloride (TC) with purity ≥ 95% was purchased from Sigma-Aldrich, Merck KGaA, Darmstadt, Germany. Nitric acid (HNO3), ZnCl2, and other analytical grade chemicals were obtained from J. T. Baker. All chemicals were used without further purification.

2.2. Adsorbent Preparation

The procedure to prepare the compound-synthetic adsorbents was similar to our recent work [28,29,30]. Orange peel (OP) was collected from a local market. After collection, the OP was washed with tap water and deionized water to remove any water-soluble impurities and adhering dirt. Then, the OP was dried at 80 °C for 24 h and crushed to 0.074–0.105 mm particles using an electric grinder. Commercial D-glucose was purchased from Sigma-Aldrich (Darmstadt, Germany). Hydrochars were prepared following a typical hydrolysis process. In brief, approximately 15 g of the precursor was added in a 150 mL Teflon-lined autoclave containing 110 mL of DI water. After a 24 h HTC process at 190 °C with a heating rate of 8 °C/min, the brown precipitate was separated by filtration. In previous studies, we have shown the results exhibited from: (i) GH vs. OPH: glucose–orange peel–hydrochar without HNO3 reflux, (ii) MGH vs. MOPH: glucose–orange peel–hydrochar with HNO3 reflux [28]. Moreover, the sawdust hydrochar samples were activated using ZnCl2 in various weight ratios (0 and 1.75) of the activating agent to the sample. The hydrochar samples were impregnated in 100 mL solutions of the activating agents in the above-mentioned weight ratios at 50 °C for 30 min. The solutions were dried in an oven at 105 °C. The dried samples were placed in the oven at 800 °C for 4 h to finish the process of activation. The products were allowed to cool to room temperature and then were washed with deionized distilled water until their pH values were approximately 7.0. The AC samples were thus synthesized: (iii) WAC: activated carbon derived from teak-sawdust hydrochars (only high-temperature calcination), and (iv) ACZ1175: AC derived from ZnCl2-activated teak-sawdust hydrochar (high-temperature, weight ratio of ZnCl2 to hydrochar = 1.75:1.0) [29,30]. Figure 1 illustrates that the preparation procedure for the hydrochar, modified-hydrochar and hydrochar-derived activated carbon by different chemical activation methods.

2.3. Adsorption Kinetics Experiment

Adsorption kinetics experiments were performed to evaluate the equilibrium time for the subsequent adsorption isotherm experiments. Approximately 0.05 g of each sample (GH, MGH, OPH, MOPH, WAC, and ACZ1175) was added to 50 mL glass tubes containing 25 mL TC solution with a concentration of 500 mg/L. All tubes were put in a table concentrator shaker and shaken at 150 rpm at 25 °C from 0.5 h to 24 h. After reaching adsorption equilibrium, the suspensions were filtered by a 0.45 µm polytetrafluoroethylene membrane filter, and aliquots of the filtrate were analyzed by UV–Vis spectroscopy (Genesys 10S UV-VIS, Thermo Scientific) at a wavelength of 355 nm (Figure 2). All experiments were conducted in duplicate, and the tubes containing only 25 mL TC solution of the same concentration were used for observing the loss of TC during this procedure. The buffer (pH = 5.5) containing acetate was used to maintain the pH of the solution during the experimental process. The amount of TC uptake at equilibrium, qe (mg/g), was calculated by the mass balance equation (Equation (1)):
q e = ( C o C e ) m 1 V 1
where Co (mg/L) and Ce (mg/L) are the TC concentrations at the beginning and equilibrium, respectively; m1 (g) is the mass of used hydrochar or AC samples; and V1 (L) is the volume of the TC solution. All batch adsorption experiments were carried out at the constant solid/liquid ratio of approximately 1.0 g/L.

2.4. Adsorption Equilibrium Experiment

Adsorption processes were performed using batch experiments. Approximately 25 mL TC aqueous solution with initial concentrations of 50–400 mg/L was placed in contact with 0.05 g adsorbent and was then shaken at 150 rpm for 24 h at room temperature. Afterwards, the following procedures, such as centrifugation, filtration, and measurement, were the same as in Section 2.3. Citrate buffer (pH 3.0), acetate buffer (pH 5.5), potassium phosphate (pH 7.0), and 2-(cyclohexyl amino) ethane-sulfonic acid (CHES-pH 9.5) buffer were used to maintain the solution pH during the test experiment.

2.5. Influence of pH Solution on TC Adsorption

The pH values of TC solution were controlled to pH 3, pH 5.5, pH 7, and pH 9.5 by using the buffer solution. After that, adsorption isotherms were conducted at room temperature (near 25 °C) in the same way as designated in Section 2.4.

2.6. Adsorption Data Analysis

In this study, the adsorption mechanisms might be evaluated based on the pseudo-first-order and pseudo-second-order models. The experimental data of the adsorption kinetics evaluated based on the linear and non-linear forms of the pseudo-first-order model [31] and pseudo-second-order model [32] are mathematically expressed in Equations (2) and (3), respectively.
q t = q e ( 1 e k 1 t )
q t = q e 2 k 2 t 1 + k 2 q e t
where k1 (1/min) and k2 (g/mg × min) are the rate constants of the pseudo-first-order and pseudo-second-order models, respectively; and qe and qt are the amounts of TC uptake per mass of the hydrochar and AC samples at equilibrium and at any time t (min), respectively.
To investigate the characterization between adsorbents and TC, the data of adsorption isotherms were fitted by Freundlich [33] and Langmuir [34] models as follows:
Freundlich   model :             q e = K F C e 1 / n F
Langmuir   model :   q e = Q m a x o K L C e 1 + K L C e
where qe is the adsorbed amount of TC in equilibrium; Ce is the equilibrium TC concentrations; Qomax (mg/g−1) is the maximum adsorption capacity of adsorbent, KL (L/mg−1) is the equilibrium constant in Langmuir equation, and KF (mg/g−1 (mg/L−1)−1/n) is the Freundlich constant. All experiments were carried out in duplicate. The average values are indicated in this study.

3. Results

3.1. Effect of pH of the TC Solution on λmax Values

The absorbance spectra of TC in various pH solutions (range of 2.11∼11.02, without adsorbent) are indicated in Figure 2. The optimal wavelength for TC in pH 2.11~9.07 solutions is found at 355 nm, while at pH > 10 the absorbance shifts to ∼385 nm. For consistency, TC aliquots of the filtrate were analyzed by UV–Vis spectroscopy at a wavelength of 355 nm (shown in Figure 2).

3.2. Morphological and Textural Properties of Synthetic Adsorbents

The partial SEM images for the synthetic adsorbents are illustrated in Figure 3. Other SEM images of adsorbent can be found in the literature adapted from Nguyen et al., 2019 and Hai et al., 2019 [28,29].
The SEM images shown in Figure 3 reveal that the sizes of these hydrochar particles range from several hundreds of nanometers to several micrometers. Interestingly, ACZ1175 showed a rugged surface with a rather large pore. The carbonaceous materials exhibited their relatively porous characteristics [28,29,35].

3.3. Characteristics of the Synthetic Adsorbents

The sizes of these hydrochar particles ranged from several hundreds of nanometers to several micrometers. In a previous study, the GH and MGH were spherical particles with a relatively more uniform size and smooth surface, whereas OPH, MOPH, and WAC displayed fragment shapes. In particular, ACZ1175 possesses an obvious large pore. The carbonaceous materials have exhibited their relatively porous characteristics in previous studies [35,36]. The detailed descriptions of characterizing the hydrochar and hydrochar-derived activated carbonaceous materials have been used in other studies [28,29].
Table 1 shows that the BET specific surface area (SBET) values of synthetic adsorbents are ranged from 4.49 to 34.06 m2/g following the increasing order MGH < GH < MOPH < OPH. The obtained result was attributed to the blocking of the narrow pores by HNO3 treatment [37]. Moreover, the increases in pore size after the oxidation process seemed to reflect this blocking effect. WAC was synthesized through high-temperature calcination without activation. Thus, WAC possessed a relatively lower SBET than the ACZ1175. Activation using ZnCl2 significantly enhanced the SBET values of the AC samples [29,38].

3.4. Adsorption Kinetic

To deeply understand the interactions of TC on hydrochar and AC samples, Figure 4 shows the results of adsorption kinetics. The qt values reached the adsorption equilibrium in around 4–5 h at room temperature. Most curves can fit the PSO model well. Furthermore, it was found that the increase in adsorption performance of MGH, MOPH, and ACZ1175 were relative to non-activated hydrochar samples.
In this study, the PFO and PSO kinetic models were used to fit the adsorption data for understanding possible mechanisms associated with the adsorption of TC on these hydrochar and hydrochar-derived ACs. Table 2 shows that the adsorption kinetic results fitted better with the PSO model than the PFO model. The R2 of PSO (0.957–0.997) were higher than the PFO model (0.807–0.977). Thus, PSO was believed to fit experimental data more accurately, similar to TC adsorption results [5,23]. Moreover, the theoretical qt,cal values calculated by the PSO kinetic model were closer to the experimental qt,exp values than fitting by the PFO kinetic model (Table 2).
Therefore, the PSO kinetic model is more appropriate to describe the kinetic adsorption of TC onto hydrochar and hydrochar-derived ACs. The adsorption behavior of TC was followed by the PSO kinetic model, which demonstrates chemical adsorption interaction between TC and these hydrochar and AC samples including forces via sharing or exchange of electrons between TC and tested adsorbents [23,24,29].

3.5. Adsorption Isotherms

Using the HNO3 modified-hydrochar and ZnCl2-hydrochar-derived AC [28,29,30,38] showed higher adsorption capacity than untreated hydrochar (Table 3); therefore, modified hydrochar and ACs were employed to explore the potential difference in adsorption between these adsorbents and TC. Adsorption data of TC onto hydrochar and AC samples at 25 °C, with different pHs, are illustrated in Figure 5.
As presented in Figure 5b, the adsorption capacities of modified hydrochar and ACs improved with increasing initial TC concentration (50–400 mg/L) at pH 5.5. To investigate the adsorption behavior between TC and all adsorbents, two kinds of isotherm models (Langmuir and Freundlich model) were used to fit the experimental data of several pH values (3.0, 5.5, 7.0, and 9.5). The parameters of TC on hydrochar and AC fitted by Langmuir and Freundlich models are listed in Table 3. Both the Langmuir and Freundlich models can fit adsorption isotherms well, as confirmed by the high R2 values (0.934–0.999), and the qe,cal is closer to qe,exp in the case of both models. The result of Langmuir model fitting indicates that the adsorption of TC on adsorbents takes place as monolayer adsorption on a surface that is homogenous in adsorption affinity. The maximum adsorption capacity of TC at 25 °C with pH 5.5 estimated by the Langmuir model was found to follow the order: ACZ1175 (257.28 mg/g) > MGH (207.11 mg/g) > WAC (197.52 mg/g) > MOPH (168.50 mg/g) > OPH (85.79 mg/g) > GH (75.47 mg/g).
Notably, both Langmuir and Freundlich models could fit the adsorption data of TC on hydrochar, modified-hydrochar and hydrochar-derived AC adsorbents well, showing the adsorption of TC onto adsorbents may be affected by multiple mechanisms. The adsorption mechanisms of TC on carbonaceous materials have been well documented in the literature, such as cation exchange, π-π interaction, H-bonding, electrostatic interaction, surface complexation, pore-filling effect, and nonspecific van der Waals force [24,25,39].
As with previous results, the HNO3 modification process enhanced oxygen- and nitrogen-containing functional groups and unsaturated bonds on the surface of modified hydrochars. The MGH has more oxygen-containing functional groups which could serve as H-bonding acceptor, and hence the adsorption of TC on MGH was higher than GH and OPH [5,28]. On the other hand, physical adsorption may contribute to the higher adsorption capacity. The ACZ1175 possessed the highest SBET (1757 m2/g−1) and pore volume (1.02 cm3/g−1) to generate high adsorption capacity although chemical adsorption dominated in this study owing to oxygen functional groups [29,38]. In this case, these results suggested that the ACZ1175 had stronger adsorption ability and higher adsorption capacity than other samples, which might be mainly explained by the higher surface area and rich oxygen-containing functional groups.

3.6. Influence of Solution pH and Adsorption Mechanism

In this study, the effects of pH on the adsorption of TC varied with the properties of TC and adsorbents. As shown in Table 4, TC is an amphoteric molecule with three pKa values (3.3; 7.68; and 9.68).
TC might alter its functional groups in different pH ranges. The predominant species are cation (H4TC+) at pH < 3.4, zwitterion (H3TCo) at 3.4 < pH < 7.6, anion (H2TC) at 7.6 < pH < 9.7, and other anion (HTC2−) at pH > 9.7 [14,40,41]. To investigate the interaction between TC and adsorbents, we used citrate buffer (pH 3.0), acetate buffer (pH 5.5), potassium phosphate (pH 7.0), and two-(cyclohexyl amino) ethanesulfonic acid (CHES—pH 9.5) buffer to maintain the solution pH during the experimental process. As with the previous results, the pHPZC values of GH, MGH, OPH, MOPH, WAC, and ACZ1175 were 6.51, 4.09, 6.25, 5.12, 8.03, and 8.11, respectively [28,29,38]. Thus, the surface charge of each adsorbent was positive charge when the solution pH was lower than their pHPZC, while it was negative at pH > pHPZC [39].
Interestingly, TC molecules (H3TCo—zwitterion species) in the pH range (3.4–7.6) carry no net electrical charge. When the hydrochar and AC have negative or positive charges on the surface, the electrostatic attraction or repulsion is minimized. Due to the MGH and ACZ1175 being rich in oxygen-containing functional groups (e.g., hydroxyl and carboxyl groups), the adsorption mechanisms of TC on the adsorbents may be ascribed to van der Waals force, surface complexation, or hydrogen bonding. The interaction of hydrogen bonding has been seldom discussed in the literature. However, the functional groups such as –CH3, –OH, –NH2, and N–H in the TC could form hydrogen bonding with –OH and –COOH on hydrochar and AC sample surfaces [23,42,43].
At higher pH > 7.0, a progressive decline in the adsorption capacities of TC on the hydrochar and AC samples was observed. This can be attributed to a highly strong electrostatic repulsion between the negatively charged TC and a negative charge on the surface of all samples in this study. The interaction here is a weak π−π electron-donor−acceptor (EDA) interaction and physical force between adsorbate and adsorbents [41]. At pH above 9.5, a decrease in the adsorption of TC is due to the higher pH solution; the surface of the sample is negatively charged. Thus, a repulsive electrostatic interaction occurred between the surface of the adsorbent and TC species (HTC2−), which decreases the TC adsorption. Moreover, another reason for the reduction is many of the reactive oxygen-containing functional groups sited on the surface of the MGH and ACZ1175, such as –OH and –COOH, were passivated or blocked when pH > 9.5.
All samples showed the highest adsorption of TC at pH 5.5 followed by pH 3 and 7, although the adsorption capacity of the raw hydrochar was lower than those of the modified-hydrochar and hydrochar-derived AC samples. In this case, the lowest adsorption capacity of TC onto the sample is at pH 9.5 (Table 3 and Figure 6).
Results of the solution pH effect revealed repulsive electrostatic interaction exists between TC and hydrochar, modified-hydrochar or hydrochar-derived AC samples. Based on the solution pH, electrostatic interaction also plays a critical role in these adsorption processes. However, electrostatic interactions were not the unique driving force for adsorption. Hence, the adsorption mechanism is affected by van der Waals force, pore filling, π-π interaction, and hydrogen bonding. The potentially different adsorbents and different mechanisms to bind contaminants onto hydrochar and hydrochar-derived ACs are shown in Figure 7.

3.7. Estimation of Adsorbent Production Cost

The adsorbents of hydrochar, modified-hydrochar, and hydrochar-derived activated carbon sourced from agricultural wastes can be used for adsorptive removal of various contaminants. The cost analysis is very important to determine whether the whole production process of the HC/AC samples is feasible or not. The production cost of adsorbent consists of various steps such as a collection of samples, size reduction, and preparation of adsorbent, carbonization, activation, and reusability [44]. The availability, treatment conditions, process requirement, and reuse are the influencing factors for cost analysis of adsorbent materials [45]. In this study, the cost estimation of preparing 1 kg adsorbent has been tabulated in Table 5.
As shown in Table 5, the cost of oxidized hydrochar (USD 4.71/kg) is slightly higher than that of hydrochar-derived activated carbon (USD 3.47/kg) due to the higher price of nitric acid which compares with ZnCl2. However, this production cost will be lower when it is produced on a large scale. The cost estimation of HC/AC evidently suggests that the activated-hydrochar preparation process using agricultural waste as a raw input is quite cost-effective. Both oxidized hydrochar and AC can be used as promising adsorbents for the removal of TC from aqueous solution.

3.8. Comparison of the Estimation of Adsorbent Production Cost

In recent years, various indigenous biomasses such as agricultural wastes have been enormously used as a sustainable precursor to produce hydrochar, biochar, and activated carbon [46]. Successful implementation of a technique for the removal of contaminants in wastewater treatment highly depends on the cost of adsorbent production [45,47].
According to the data, the costs of adsorbents derived from different biomasses have been calculated and summarized as being in a range of USD 2–6/kg (Table 6), excepting commercial products. From this comparative analysis, modified-hydrochar and hydrochar-derived AC as adsorbents are inexpensive or even have negligible cost. The production technique is also very inexpensive and does not consume costly reagents, in comparison to AC production and activation, which require high temperature and expensive reagents.

4. Conclusions

The objective of this study was to analyze and compare the behavior of carbonaceous materials, both hydrochar and hydrochar-derived activated carbon, with different chemical and textural natures in the adsorption of TC. The following conclusions can be presented in this study:
In order to improve the adsorption capacity of TC, HNO3 and ZnCl2 were applied under different conditions to activate hydrochars sourced from agricultural waste. The maximum adsorption capacities of TC estimated by the Langmuir model were found to follow the order: ACZ1175 (257.28 mg/g) > MGH (207.11 mg/g) > WAC (197.52 mg/g) > MOPH (168.50 mg/g) > OPH (85.79 mg/g) > GH (75.47 mg/g) in pH 5.5 solution at 25 °C. Modified-hydrochar and hydrochar-derived AC adsorbents sourced from agricultural waste showed high efficiency for removal of TC in wastewater.
The adsorption kinetic curves could fit the PSO model well. Chemical adsorption is regarded as the primary adsorption mechanism. The solution pH and ionic strength had a major influence on TC adsorption on the carbons, indicating that electrostatic adsorbent–adsorbate interactions also play an important role in TC adsorption at pH values that produce TC deprotonation. In addition, the adsorption mechanisms of TC on the test adsorbents also included hydrogen bonding, π-π interaction, pore-filling, and electrostatic interactions.
Most significantly, the total production cost spent on the preparation of hydrochar-derived activated carbon is a little bit cheaper as compared with oxidized hydrochar. The good cost made it a feasible adsorbent instead of commercial-activated carbon for potentially wide application in wastewater treatment.

Author Contributions

Conceptualization and experimental design, D.M.N., H.H.C., N.H.T. and N.Q.T.; writing—original draft preparation, D.M.N.; writing—review and editing, D.M.N. and N.C.H.; visualization and supervision, N.C.H., N.D.H. and H.-P.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the Faculty of Environment, Thai Nguyen University of Agriculture and Forestry (TUAF).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data that support the findings of this study are available from the corresponding author, [N.D.H. and H.-P.C.], upon reasonable request.

Acknowledgments

We thank all of our laboratory colleagues for their constant support. We also thank the comments and suggestions from the reviewers, which greatly improved this manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

WACActivated carbon derived from teak-sawdust hydrochars (only high-temperature calcination without activation process)
ACZ1175AC derived from ZnCl2-activated teak-sawdust hydrochar (high-temperature, weight ratio of ZnCl2 to hydrochar = 1.75:1.0)
GHHydrochars derived from glucose solution through a hydrothermal process
OPHHydrochars derived from orange peel through a hydrothermal process
MGHHydrochars derived from glucose solution through a hydrothermal process, and then were activated by HNO3
MOPHHydrochars derived from orange peel through a hydrothermal process, and then were activated by HNO3

References

  1. Chen, Z.; Zhang, W.; Wang, G.; Zhang, Y.; Gao, Y.; Boyd, S.A.; Teppen, B.J.; Tiedje, J.M.; Zhu, D.; Li, H. Bioavailability of soil-sorbed tetracycline to Escherichia coli under unsaturated conditions. Environ. Sci. Technol. 2017, 51, 6165–6173. [Google Scholar] [CrossRef] [PubMed]
  2. Jing, X.-R.; Wang, Y.-Y.; Liu, W.-J.; Wang, Y.-K.; Jiang, H. Enhanced adsorption performance of tetracycline in aqueous solutions by methanol-modified biochar. Chem. Eng. J. 2014, 248, 168–174. [Google Scholar] [CrossRef]
  3. Bao, Y.; Zhou, Q.; Wan, Y.; Yu, Q.; Xie, X. Effects of soil/solution ratios and cation types on adsorption and desorption of tetracycline in soils. Soil Sci. Soc. Am. J. 2010, 74, 1553–1561. [Google Scholar] [CrossRef] [Green Version]
  4. Lian, F.; Song, Z.; Liu, Z.; Zhu, L.; Xing, B. Mechanistic understanding of tetracycline sorption on waste tire powder and its chars as affected by Cu2+ and pH. Environ. Pollut. 2013, 178, 264–270. [Google Scholar] [CrossRef] [PubMed]
  5. Chen, T.; Luo, L.; Deng, S.; Shi, G.; Zhang, S.; Zhang, Y.; Deng, O.; Wang, L.; Zhang, J.; Wei, L. Sorption of tetracycline on H3PO4 modified biochar derived from rice straw and swine manure. Bioresour. Technol. 2018, 267, 431–437. [Google Scholar] [CrossRef] [PubMed]
  6. Chen, Y.-Y.; Ma, Y.-L.; Yang, J.; Wang, L.-Q.; Lv, J.-M.; Ren, C.-J. Aqueous tetracycline degradation by H2O2 alone: Removal and transformation pathway. Chem. Eng. J. 2017, 307, 15–23. [Google Scholar] [CrossRef]
  7. Guan, R.; Yuan, X.; Wu, Z.; Jiang, L.; Zhang, J.; Li, Y.; Zeng, G.; Mo, D. Efficient degradation of tetracycline by heterogeneous cobalt oxide/cerium oxide composites mediated with persulfate. Sep. Purif. Technol. 2019, 212, 223–232. [Google Scholar] [CrossRef]
  8. Wang, W.; Fang, J.; Shao, S.; Lai, M.; Lu, C. Compact and uniform TiO2@g-C3N4 core-shell quantum heterojunction for photocatalytic degradation of tetracycline antibiotics. Appl. Catal. B Environ. 2017, 217, 57–64. [Google Scholar] [CrossRef]
  9. Yan, M.; Hua, Y.; Zhu, F.; Gu, W.; Jiang, J.; Shen, H.; Shi, W. Fabrication of nitrogen doped graphene quantum dots-BiOI/MnNb2O6 pn junction photocatalysts with enhanced visible light efficiency in photocatalytic degradation of antibiotics. Appl. Catal. B Environ. 2017, 202, 518–527. [Google Scholar] [CrossRef]
  10. Leng, Y.; Bao, J.; Chang, G.; Zheng, H.; Li, X.; Du, J.; Snow, D.; Li, X. Biotransformation of tetracycline by a novel bacterial strain Stenotrophomonas maltophilia DT1. J. Hazard. Mater. 2016, 318, 125–133. [Google Scholar] [CrossRef] [Green Version]
  11. Sousa, J.M.; Macedo, G.; Pedrosa, M.; Becerra-Castro, C.; Castro-Silva, S.; Pereira, M.F.R.; Silva, A.M.; Nunes, O.C.; Manaia, C.M. Ozonation and UV254 nm radiation for the removal of microorganisms and antibiotic resistance genes from urban wastewater. J. Hazard. Mater. 2017, 323, 434–441. [Google Scholar] [CrossRef]
  12. Chen, H.; Zhang, M. Occurrence and removal of antibiotic resistance genes in municipal wastewater and rural domestic sewage treatment systems in eastern China. Environ. Int. 2013, 55, 9–14. [Google Scholar] [CrossRef]
  13. Liu, M.; Hou, L.-a.; Yu, S.; Xi, B.; Zhao, Y.; Xia, X. MCM-41 impregnated with A zeolite precursor: Synthesis, characterization and tetracycline antibiotics removal from aqueous solution. Chem. Eng. J. 2013, 223, 678–687. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Jang, H.M.; Yoo, S.; Choi, Y.-K.; Park, S.; Kan, E. Adsorption isotherm, kinetic modeling and mechanism of tetracycline on Pinus taeda-derived activated biochar. Bioresour. Technol. 2018, 259, 24–31. [Google Scholar] [CrossRef] [PubMed]
  15. Zeng, Z.; Ye, S.; Wu, H.; Xiao, R.; Zeng, G.; Liang, J.; Zhang, C.; Yu, J.; Fang, Y.; Song, B. Research on the sustainable efficacy of g-MoS2 decorated biochar nanocomposites for removing tetracycline hydrochloride from antibiotic-polluted aqueous solution. Sci. Total Environ. 2019, 648, 206–217. [Google Scholar] [CrossRef] [PubMed]
  16. Gao, Y.; Li, Y.; Zhang, L.; Huang, H.; Hu, J.; Shah, S.M.; Su, X. Adsorption and removal of tetracycline antibiotics from aqueous solution by graphene oxide. J. Colloid Interface Sci. 2012, 368, 540–546. [Google Scholar] [CrossRef]
  17. Chen, W.-R.; Huang, C.-H. Adsorption and transformation of tetracycline antibiotics with aluminum oxide. Chemosphere 2010, 79, 779–785. [Google Scholar] [CrossRef]
  18. Liu, H.; Yang, Y.; Kang, J.; Fan, M.; Qu, J. Removal of tetracycline from water by Fe-Mn binary oxide. J. Environ. Sci. 2012, 24, 242–247. [Google Scholar] [CrossRef]
  19. Parolo, M.E.; Avena, M.J.; Pettinari, G.R.; Baschini, M.T. Influence of Ca2+ on tetracycline adsorption on montmorillonite. J. Colloid Interface Sci. 2012, 368, 420–426. [Google Scholar] [CrossRef] [PubMed]
  20. Ahsan, M.A.; Islam, M.T.; Hernandez, C.; Castro, E.; Katla, S.K.; Kim, H.; Lin, Y.; Curry, M.L.; Gardea-Torresdey, J.; Noveron, J.C. Biomass conversion of saw dust to a functionalized carbonaceous materials for the removal of Tetracycline, Sulfamethoxazole and Bisphenol A from water. J. Environ. Chem. Eng. 2018, 6, 4329–4338. [Google Scholar] [CrossRef]
  21. Álvarez-Torrellas, S.; Rodríguez, A.; Ovejero, G.; García, J. Comparative adsorption performance of ibuprofen and tetracycline from aqueous solution by carbonaceous materials. Chem. Eng. J. 2016, 283, 936–947. [Google Scholar] [CrossRef]
  22. Rivera-Utrilla, J.; Gómez-Pacheco, C.V.; Sánchez-Polo, M.; López-Peñalver, J.J.; Ocampo-Pérez, R. Tetracycline removal from water by adsorption/bioadsorption on activated carbons and sludge-derived adsorbents. J. Environ. Manag. 2013, 131, 16–24. [Google Scholar] [CrossRef] [PubMed]
  23. Sayğılı, H.; Güzel, F. Effective removal of tetracycline from aqueous solution using activated carbon prepared from tomato (Lycopersicon esculentum Mill.) industrial processing waste. Ecotoxicol. Environ. Saf. 2016, 131, 22–29. [Google Scholar] [CrossRef] [PubMed]
  24. Peiris, C.; Gunatilake, S.R.; Mlsna, T.E.; Mohan, D.; Vithanage, M. Biochar based removal of antibiotic sulfonamides and tetracyclines in aquatic environments: A critical review. Bioresour. Technol. 2017, 246, 150–159. [Google Scholar] [CrossRef] [PubMed]
  25. Ji, L.; Wan, Y.; Zheng, S.; Zhu, D. Adsorption of tetracycline and sulfamethoxazole on crop residue-derived ashes: Implication for the relative importance of black carbon to soil sorption. Environ. Sci. Technol. 2011, 45, 5580–5586. [Google Scholar] [CrossRef]
  26. Liu, P.; Liu, W.-J.; Jiang, H.; Chen, J.-J.; Li, W.-W.; Yu, H.-Q. Modification of bio-char derived from fast pyrolysis of biomass and its application in removal of tetracycline from aqueous solution. Bioresour. Technol. 2012, 121, 235–240. [Google Scholar] [CrossRef]
  27. Zhu, X.; Liu, Y.; Zhou, C.; Luo, G.; Zhang, S.; Chen, J. A novel porous carbon derived from hydrothermal carbon for efficient adsorption of tetracycline. Carbon 2014, 77, 627–636. [Google Scholar] [CrossRef]
  28. Nguyen, D.H.; Tran, H.N.; Chao, H.-P.; Lin, C.-C. Effect of nitric acid oxidation on the surface of hydrochars to sorb methylene blue: An adsorption mechanism comparison. Adsorpt. Sci. Technol. 2019, 37, 607–622. [Google Scholar] [CrossRef] [Green Version]
  29. Duy Nguyen, H.; Nguyen Tran, H.; Chao, H.-P.; Lin, C.-C. Activated carbons derived from teak sawdust-hydrochars for efficient removal of methylene blue, copper, and cadmium from aqueous solution. Water 2019, 11, 2581. [Google Scholar] [CrossRef] [Green Version]
  30. Tran, H.N.; Huang, F.-C.; Lee, C.-K.; Chao, H.-P. Activated carbon derived from spherical hydrochar functionalized with triethylenetetramine: Synthesis, characterizations, and adsorption application. Green Process. Synth. 2017, 6, 565–576. [Google Scholar] [CrossRef]
  31. Corbett, J.F. Pseudo first-order kinetics. J. Chem. Educ. 1972, 49, 663. [Google Scholar] [CrossRef]
  32. Ho, Y.-S.; McKay, G. Pseudo-second order model for sorption processes. Process Biochem. 1999, 34, 451–465. [Google Scholar] [CrossRef]
  33. Freundlich, H. Uber die adsorption in losungen. Z. Fur Phys. Chem.-Leipz. 1907, 57, 385–470. [Google Scholar] [CrossRef]
  34. Langmuir, I. The adsorption of gases on plane surfaces of glass, mica and platinum. J. Am. Chem. Soc. 1918, 40, 1361–1403. [Google Scholar] [CrossRef] [Green Version]
  35. Tran, H.N.; You, S.-J.; Chao, H.-P. Insight into adsorption mechanism of cationic dye onto agricultural residues-derived hydrochars: Negligible role of π-π interaction. Korean J. Chem. Eng. 2017, 34, 1708–1720. [Google Scholar] [CrossRef]
  36. Sevilla, M.; Fuertes, A.B. Chemical and structural properties of carbonaceous products obtained by hydrothermal carbonization of saccharides. Chem. Eur. J. 2009, 15, 4195–4203. [Google Scholar] [CrossRef]
  37. Boehm, H.; Tereozki, B.; Schanz, K. Blocking of pores in porous carbons by chemisorption. Stud. Surf. Sci. Catal. 1982, 10, 395–401. [Google Scholar]
  38. Vo, A.T.; Nguyen, V.P.; Ouakouak, A.; Nieva, A.; Doma, B.T., Jr.; Tran, H.N.; Chao, H.-P. Efficient removal of Cr (VI) from water by biochar and activated carbon prepared through hydrothermal carbonization and pyrolysis: Adsorption-coupled reduction mechanism. Water 2019, 11, 1164. [Google Scholar] [CrossRef] [Green Version]
  39. Ji, L.; Chen, W.; Duan, L.; Zhu, D. Mechanisms for strong adsorption of tetracycline to carbon nanotubes: A comparative study using activated carbon and graphite as adsorbents. Environ. Sci. Technol. 2009, 43, 2322–2327. [Google Scholar] [CrossRef]
  40. Zhu, X.; Liu, Y.; Qian, F.; Zhou, C.; Zhang, S.; Chen, J. Preparation of magnetic porous carbon from waste hydrochar by simultaneous activation and magnetization for tetracycline removal. Bioresour. Technol. 2014, 154, 209–214. [Google Scholar] [CrossRef]
  41. Zhou, Y.; Liu, X.; Xiang, Y.; Wang, P.; Zhang, J.; Zhang, F.; Wei, J.; Luo, L.; Lei, M.; Tang, L. Modification of biochar derived from sawdust and its application in removal of tetracycline and copper from aqueous solution: Adsorption mechanism and modelling. Bioresour. Technol. 2017, 245, 266–273. [Google Scholar] [CrossRef] [PubMed]
  42. Zhang, G.; Liu, X.; Sun, K.; He, F.; Zhao, Y.; Lin, C. Competitive Sorption of Metsulfuron-Methyl and Tetracycline on Corn Straw Biochars. J. Environ. Qual. 2012, 41, 1906–1915. [Google Scholar] [CrossRef] [PubMed]
  43. Yang, Q.; Wu, P.; Liu, J.; Rehman, S.; Ahmed, Z.; Ruan, B.; Zhu, N. Batch interaction of emerging tetracycline contaminant with novel phosphoric acid activated corn straw porous carbon: Adsorption rate and nature of mechanism. Environ. Res. 2020, 181, 108899. [Google Scholar] [CrossRef]
  44. Banerjee, M.; Basu, R.K.; Das, S.K. Cr (VI) adsorption by a green adsorbent walnut shell: Adsorption studies, regeneration studies, scale-up design and economic feasibility. Process Saf. Environ. Prot. 2018, 116, 693–702. [Google Scholar] [CrossRef]
  45. Banerjee, S.; Barman, S.; Halder, G. Sorptive elucidation of rice husk ash derived synthetic zeolite towards deionization of coalmine waste water: A comparative study. Groundw. Sustain. Dev. 2017, 5, 137–151. [Google Scholar] [CrossRef]
  46. Zheng, Y.; Tao, L.; Yang, X.; Huang, Y.; Liu, C.; Zheng, Z. Study of the thermal behavior, kinetics, and product characterization of biomass and low-density polyethylene co-pyrolysis by thermogravimetric analysis and pyrolysis-GC/MS. J. Anal. Appl. Pyrolysis 2018, 133, 185–197. [Google Scholar] [CrossRef]
  47. Chakraborty, P.; Show, S.; Banerjee, S.; Halder, G. Mechanistic insight into sorptive elimination of ibuprofen employing bi-directional activated biochar from sugarcane bagasse: Performance evaluation and cost estimation. J. Environ. Chem. Eng. 2018, 6, 5287–5300. [Google Scholar] [CrossRef]
  48. Yihunu, E.W.; Minale, M.; Abebe, S.; Limin, M. Preparation, characterization and cost analysis of activated biochar and hydrochar derived from agricultural waste: A comparative study. SN Appl. Sci. 2019, 1, 1–8. [Google Scholar] [CrossRef] [Green Version]
  49. Mondal, S.; Aikat, K.; Halder, G. Biosorptive uptake of ibuprofen by chemically modified Parthenium hysterophorus derived biochar: Equilibrium, kinetics, thermodynamics and modeling. Ecol. Eng. 2016, 92, 158–172. [Google Scholar] [CrossRef]
  50. Wang, Y.; Huang, C.; Xu, T. Which is more competitive for production of organic acids, ion-exchange or electrodialysis with bipolar membranes? J. Membr. Sci. 2011, 374, 150–156. [Google Scholar] [CrossRef]
Figure 1. Schematic process for the preparation of hydrochar (GH, OPH), modification of hydrochar (MGH, MOPH) and hydrochar-derived activated carbon (WAC, and ACZ1175).
Figure 1. Schematic process for the preparation of hydrochar (GH, OPH), modification of hydrochar (MGH, MOPH) and hydrochar-derived activated carbon (WAC, and ACZ1175).
Sustainability 15 04412 g001
Figure 2. Effect of pH of the TC solution on λmax values (without adsorbent).
Figure 2. Effect of pH of the TC solution on λmax values (without adsorbent).
Sustainability 15 04412 g002
Figure 3. SEM images of WAC (a) and ACZ1175 (b) samples.
Figure 3. SEM images of WAC (a) and ACZ1175 (b) samples.
Sustainability 15 04412 g003
Figure 4. Adsorption kinetics of TC with the concentration of 500 mg/L−1 on the hydrochar, modified-hydrochar, and hydrochar-derived AC samples at 25 °C.
Figure 4. Adsorption kinetics of TC with the concentration of 500 mg/L−1 on the hydrochar, modified-hydrochar, and hydrochar-derived AC samples at 25 °C.
Sustainability 15 04412 g004
Figure 5. Adsorption isotherm of TC on the hydrochar, modified-hydrochar, and hydrochar-derived AC samples with (a) pH 3.0; (b) pH 5.5; (c) pH 7.0; (d) pH 9.5, at 25 °C.
Figure 5. Adsorption isotherm of TC on the hydrochar, modified-hydrochar, and hydrochar-derived AC samples with (a) pH 3.0; (b) pH 5.5; (c) pH 7.0; (d) pH 9.5, at 25 °C.
Sustainability 15 04412 g005
Figure 6. Effect of pH on the adsorption of TC on hydrochar, modified-hydrochar, and hydrochar-derived AC samples. Conditions: temperature 25 °C; [TC] = 50 mg/L−1; adsorbent dosage 0.05 g; different pH values (3.0; 5.5; 7.0; and 9.5).
Figure 6. Effect of pH on the adsorption of TC on hydrochar, modified-hydrochar, and hydrochar-derived AC samples. Conditions: temperature 25 °C; [TC] = 50 mg/L−1; adsorbent dosage 0.05 g; different pH values (3.0; 5.5; 7.0; and 9.5).
Sustainability 15 04412 g006
Figure 7. Proposed mechanism for TC adsorption by hydrochar, modified−hydrochar, and hydrochar−derived ACs.
Figure 7. Proposed mechanism for TC adsorption by hydrochar, modified−hydrochar, and hydrochar−derived ACs.
Sustainability 15 04412 g007
Table 1. Textural properties of the synthetic adsorbents.
Table 1. Textural properties of the synthetic adsorbents.
Name of SamplesSBET
(m2/g)
Pore Volume
(cm3/g)
GH7.330.015
MGH4.490.012
OPH34.060.047
MOPH20.250.042
WAC7920.340
ACZ117517571.020
Table 2. Adsorption kinetics parameters of TC on the hydrochar and AC samples.
Table 2. Adsorption kinetics parameters of TC on the hydrochar and AC samples.
Kinetic ModelsPseudo-First-OrderPseudo-Second-Order
Sampleqt,expqt,cal = qt(1 − e−k1t)qt,cal = (k2 qt2t)/(1 + k2 qt2t)
qt,calk1R2qt,calk2R2
GH57.4250.640.430.89957.490.0090.957
MGH194.83184.190.640.977198.780.0300.997
OPH67.9653.890.540.91860.950.0110.974
MOPH154.74135.210.880.912144.140.0210.972
WAC180.19167.170.910.965176.630.0350.997
ACZ1175243.08205.890.940.807234.290.0370.992
Table 3. The isotherm parameters for the adsorption of TC onto the hydrochar, modified-hydrochar and hydrochar-derived AC samples.
Table 3. The isotherm parameters for the adsorption of TC onto the hydrochar, modified-hydrochar and hydrochar-derived AC samples.
Temperature/Samples Langmuir ParametersFreundlich Parameters
qe = (QomaxKLCe)/(1 + KLCe)qe = KFCe1/nF
Qe,exp (mg/g)Qo(max)
(mg/g)
KL
(L/mg)
R2KF
(mg/g)
nFR2
Initial pH = 3
GH42.4058.460.0050.9901.950.510.967
MGH137.25134.980.0150.9849.920.440.939
OPH57.6365.090.0060.9912.230.520.969
MOPH85.5781.080.0110.9935.420.430.962
WAC97.97101.090.0140.9948.340.410.965
ACZ1175160.74178.320.0600.99328.290.370.972
Initial pH = 5.5
GH58.3775.470.0060.9872.940.520.963
MGH195.22207.110.0940.97543.350.330.983
OPH67.9685.790.0100.9985.380.420.974
MOPH154.27168.500.0110.9967.210.530.979
WAC180.19197.520.0260.99615.010.490.995
ACZ1175243.08257.280.2990.99964.240.500.981
Initial pH = 7
GH55.0269.420.0050.9841.970.560.956
MGH157.98172.770.0140.98510.560.480.943
OPH65.5481.680.0090.9944.470.430.964
MOPH126.54135.540.0110.9767.640.480.934
WAC139.65149.040.0170.98611.110.450.946
ACZ1175181.11190.860.1160.98239.980.360.972
Initial pH = 9.5
GH40.6551.210.0020.9680.330.730.951
MGH115.51126.230.0100.9955.310.510.979
OPH51.9755.660.0050.9951.270.550.978
MOPH58.6961.110.0110.9934.660.410.997
WAC70.2173.200.0130.9939.190.340.956
ACZ1175123.62157.940.0170.9988.510.490.979
Table 4. Examples of the chemical formula and characteristics of tetracycline.
Table 4. Examples of the chemical formula and characteristics of tetracycline.
Property NameChemical FormulaMolecular WeightMolecular StructurepKa1pKa2pKa3
Tetracycline (TC)C22H25ClN2O8480.9 g/molSustainability 15 04412 i0013.307.689.68
Table 5. Cost estimation of compound-synthetic adsorbent production.
Table 5. Cost estimation of compound-synthetic adsorbent production.
(A) Cost Estimation of Modified Hydrochar
ParticularsSub-SectionsCost Break-UpTotal Cost (USD)
Processing of raw materialCollection of raw materialOrange peel was collected free of cost from the local market 0.0
Hand sorting and washing DI water gained from laboratory set-up0.0
Drying costIt was done under the sun0.0
Preparation of oxidized hydrocharHydrothermal carbonization costHour × unit × cost per unit = 24 × 0.5 × 0.161.92
Acid reflux with HNO3Unit consumed × cost per unit = 13.9 mL × 78.7/500 mL2.18
Cost of dryingHour × unit × cost per unit = 6 × 0.3 × 0.16 0.28
Net cost4.38
10% to overhead charge0.43
Total cost4.71
(B) Cost Estimation of Hydrochar-Derived Activated Carbon
Processing of raw materialCollection of raw materialThe teak (T. grandis) sawdust was obtained from a furniture factory0.0
Washing costDI water gained from laboratory set-up0.0
Drying costIt was done under the sun0.0
Preparation of AC Hydrothermal carbonization costHour × unit × cost per unit = 24 × 0.5 × 0.161.92
Impregnation with ZnCl2Unit consumed × cost per unit = 1.75 g × 94/500 g0.32
Pyrolysis at 800 °CHour × unit × cost per unit = 4 × 1 × 0.160.64
Washing costDI water gained from laboratory set-up0.0
Cost of dryingHour × unit × cost per unit = 6 × 0.3 × 0.16 0.28
Net cost3.16
10% to overhead charge0.31
Total cost3.47
Table 6. Comparative analysis with other biomass-based chars reported in the literature.
Table 6. Comparative analysis with other biomass-based chars reported in the literature.
Raw MaterialsPre–Post TreatmentPost-TreatmentSBET Surface Area (m2/g)Cost/kg (USD)References
Sugarcane bagassePyrolysisH3PO45573.81[47]
Super-heated steam3.49
Orange peelHTCHNO3204.71This study
Teak sawdustHTCZnCl217573.47This study
Teff strawPyrolysisH3PO46273.73[48]
HTCH3PO4432.93
Rice husk Pyrolysis Zeolite
Z-RHA
765.42[45]
Parthenium hysterophorusPyrolysisNaOH3082.88[49]
Ion-exchange resins ---150[50]
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

Ngoc, D.M.; Hieu, N.C.; Trung, N.H.; Chien, H.H.; Thi, N.Q.; Hai, N.D.; Chao, H.-P. Tetracycline Removal from Water by Adsorption on Hydrochar and Hydrochar-Derived Activated Carbon: Performance, Mechanism, and Cost Calculation. Sustainability 2023, 15, 4412. https://doi.org/10.3390/su15054412

AMA Style

Ngoc DM, Hieu NC, Trung NH, Chien HH, Thi NQ, Hai ND, Chao H-P. Tetracycline Removal from Water by Adsorption on Hydrochar and Hydrochar-Derived Activated Carbon: Performance, Mechanism, and Cost Calculation. Sustainability. 2023; 15(5):4412. https://doi.org/10.3390/su15054412

Chicago/Turabian Style

Ngoc, Duong Minh, Nguyen Chi Hieu, Nguyen Huy Trung, Hoang Huu Chien, Nguyen Quang Thi, Nguyen Duy Hai, and Huan-Ping Chao. 2023. "Tetracycline Removal from Water by Adsorption on Hydrochar and Hydrochar-Derived Activated Carbon: Performance, Mechanism, and Cost Calculation" Sustainability 15, no. 5: 4412. https://doi.org/10.3390/su15054412

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