Next Article in Journal
A Framework for Water Security Data Gathering Strategies
Next Article in Special Issue
Low-Cost Adsorbents for the Removal of Phenol/Phenolics, Pesticides, and Dyes from Wastewater Systems: A Review
Previous Article in Journal
Developing a Decision Support Evaluation Model Based on the Matter Element Analysis Method to Optimize the Environmental Flows in Dammed Rivers
Previous Article in Special Issue
Enhancing the Efficiency of Banana Peel Bio-Coagulant in Turbid and River Water Treatment Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Selective and Competitive Adsorption of Anions in Solution on Porous Adsorbent from Zea mays Steams: Kinetic and Equilibrium Study

1
Process Design and Biomass Utilization Research Group (IDAB), Chemical Engineering Department, Engineering Faculty, Universidad de Cartagena, Avenida del Consulado St. 30, Cartagena de Indias 130015, Colombia
2
Food Packaging and Shelf Life Research Group (FP&SL), Food Engineering Department, Engineering Faculty Universidad de Cartagena, Avenida del Consulado St. 30, Cartagena de Indias 130015, Colombia
*
Authors to whom correspondence should be addressed.
Water 2022, 14(18), 2906; https://doi.org/10.3390/w14182906
Submission received: 6 July 2022 / Revised: 7 September 2022 / Accepted: 15 September 2022 / Published: 17 September 2022
(This article belongs to the Special Issue Solid/Liquid Adsorption in Water and Wastewater Treatment)

Abstract

:
Surface modification can improve the adsorption capacity of biochar. Biochar was produced from corn stalks (Zea mays) by pyrolysis at 520 °C, activated with sulfuric acid (H2SO4) using impregnation ratios of biomass weight: volume 1:2 (B 1:2) and 1:3 (B 1:3). The kinetic study showed that the equilibrium is reached at 180 min; the maximum adsorption capacity of nitrate and sulphate was obtained with B 1:2 and for phosphate with B 1:3. The adsorption of nitrate and phosphate with the two biochars presented a good fitting to the pseudo-first order and pseudo-second-order model, while that of sulphate for B 1:2 is described by Elovich’s model. Freundlich’s model describes the equilibrium of adsorption of nitrate and phosphate using B 1:2 and B 1:3, while Dubinin–Radushkevich adjusts the removal of sulphate for C 1:2 and C 1:3; therefore it is suggested that adsorption occurs in multilayers. The multicomponent study evidences the preference of biocarbon for phosphate, without indicating competition for the active centers of the material among the anions studied.

1. Introduction

Significant use and excessive discharge of phosphorus, nitrogen, and sulphur compounds increase the concentration of nutrients in surface and groundwater bodies [1]. Common sources that generate contamination from the discharge of nitrate, sulphate, and phosphate into water bodies generally arise from the waste products of human activities, such as discharges from industrialized practices, domestic waters, agricultural uses such as inorganic fertilizers, compost, and wastewater treatment effluents [2]. The excessive concentration of nutrients in the water accelerates the growth of blue-green algae and causes the death of fish due to the reduction in available oxygen [3]. Therefore, the minimum concentration of phosphate, nitrate, and sulphate capable of inducing eutrophication was set at 0.02 mg/L [4,5].
On the other hand, the intake of these nutrients by humans can cause diarrhea, miscarriages, respiratory tract infections, vomiting, hypertension, and methemoglobinemia. Therefore, the US-EPA and the World Health Organization (WHO) establish in drinking water concentrations of 10 µg/L of phosphorus [6], 10 mg/L of nitrate [7], and 250 mg/L of sulphate [5]. For this reason, it is necessary to control the concentrations of nitrate, sulphate, and phosphate in effluents before discharging them into aquatic environments. Thus, several physicochemical methods have been used in the removal of these pollutants in water, including chemical precipitation [8], oxidation [9], artificial wetlands [10], electrodialysis [11], membrane processes [12], and adsorption [13,14]. Adsorption proved to be one of the most popular techniques among those mentioned, due to its low cost, ease of operation, higher disposal efficiency, and excellent reuse performance [15]. The selection of adsorbent material during the adsorption process is a key parameter; therefore, research has been conducted to develop sustainable and economical adsorbents to improve the remediation performance of water stream contaminants [16].
Among the bioadsorbents studied, biochar is one of the most popular due to its good performance; it is a solid carbon-rich by-product of biomass pyrolysis under the complete exclusion of oxygen at temperatures below 700 °C [17], can be derived from a wide range of low-cost biomass sources, including post-harvest agricultural residues, manure, bioenergy crops, residual sludge, and organic waste [1]. It has been reported that biochar has a high density of surface charge to increase ion exchange capacity, with a large specific surface area, internal porosity, and a polar and non-polar functional group at the surface to absorb nutrients [18]. Several studies have used biochar as an adsorbent to remove, individually or simultaneously, nutrients present in water [17,19,20,21], reporting that removal efficiencies vary depending on the types and properties of the adsorbents and the environmental conditions of the water phase.
Thus, this research aims to study the kinetics and equilibrium of adsorption of nitrate (NO3), sulphate (SO4), and phosphate (PO43) anions on porous carbon-type adsorbent from sulphuric acid-modified Zea mays (H2SO4). The effect of the impregnation ratio was investigated and the effect of competition between anions was determined.

2. Materials and Methods

2.1. Materials

To develop the kinetic and equilibrium study, synthetic solutions were prepared at 100 mg/L using distilled water and monopotassium phosphate (KH2PO4), sodium nitrate (NaNO3), and potassium sulphate (K2SO4) of analytical grade, Merk Millipore with 95% purity. The biochar was activated with sulfuric acid (H2SO4). The concentration of the anions was determined using a Biobase UV/VIS Spectrophotometer model BK-UV1900.

2.2. Preparation of Bioadsorbents

To prepare the biochar, post-harvest corn stalks were used as raw material, gathered in a village in Bolivar, Colombia. The raw material was salted with deionized water, dried for 12 h at 60 °C and the size reduction was done in an electric mill. The particle size was selected between 1–2 mm. The biomass was impregnated with a 50% solution of H2SO4 with impregnation ratios 1:2 (B 1:2) and 1:3 p/v (B 1:3); then it was carbonized in a muffle at 520 °C for 30 min with a 10 °C/min ramp. The biochar was cooled down to room temperature, washed with distilled water to neutral pH, dried at 100 °C to constant mass, and stored in airtight containers until its use in the adsorption tests [22]. The prepared materials were used previously in the evaluation of the effect of temperature [23].

2.3. Adsorption Kinetics

Adsorption kinetics provides information about the service time of the adsorbent and possible mechanisms of adsorption and to describe adsorption-adsorbent interactions. Kinetic experiments were conducted at 25 °C, at 100 mg/L using 0.2 g in 100 mL solution, 200 rpm sampling at 5, 10, 20, 30, 60, 180, 300, 420, and 1440 min for each experiment. The final concentration of phosphate nitrate, and sulphate in the solution was determined by UV-Vis spectrophotometry at 880 nm [24], 543 nm [25], and 420 nm, respectively [26]. The adsorption capacity was determined according to Equation (1), where qt is the adsorption capacity at time t, Ci and Cf are the initial and final concentration of pollutant, respectively, in mg L−1, V is the volume of the solution in L and m the mass of the adsorbent in g:
q t m g g = C i C f × V m
The kinetic experiments indicate the duration needed to reach the equilibrium condition, therefore the models of pseudo-first order (Equation (2)), pseudo-second order (Equation (3)), and Elovich (Equation (4)) were adjusted to analyze the kinetic experimental data using non-linear regression due to a decrease in errors in OriginPro® (Northampton, MA, USA), with the sum of errors (SS) as adjustment criteria.
The pseudo-first-order Lagergren model describes liquid phase adsorption processes, where it is assumed that the reaction rate is driven by only one of the reagents [27]. Where k1 (min−1) is the Lagergren’s constant, qe is the amount of contaminant adsorbed per unit mass in the equilibrium, qt is the amount of contaminant adsorbed per unit of mass at any time t, and t is the time.
q t = q e 1 + e k t
The pseudo-second equation assumes that the adsorption phenomenon occurs due to the ion exchange produced at the surface of the adsorbent and that the order of the reaction is two with respect to the number of adsorption sites available for exchange [28]. Where k2 (g mmol−1 min−1) is the second-order reaction rate coefficient, t is the time in minutes, qe the equilibrium adsorption capacity (mmol g−1), and qt is the adsorption capacity at time t.
q t = t 1 k 2 × q e 2 + t q e
The Elovich equation is commonly used to describe the chemisorption of gases on solids; however, it is also used in the kinetics of adsorption of aqueous solutions on solid surfaces [29]. Where qt is the adsorption capacity at time t, α (mg g−1 min−1) is the initial adsorption rate, and β (g mg−1) is the desorption constant related to surface range and activation energy for chemisorption.
q t = q β × ln α β + 1 β × ln t

2.4. Adsorption Isotherms

Adsorption isotherms are the graphic representation of the amount of solute adsorbed per unit mass of adsorbent, describing the phenomena responsible for the process. Its models were developed based on the thermodynamic equilibrium of adsorption [30]. The experiments were performed using different concentrations of the anions: 20, 40, 60, 80, and 100 mg/L, for 24 h, with 100 mL of solution and 2 g/L of adsorbent dose at 25 °C. The models of Langmuir (Equation (5)), Freundlich (Equation (6)), and Dubinin–Radushkevich (Equation (7)) were used to adjust to the experimental data obtained at equilibrium time.
Langmuir’s isothermal model is widely used for the description of adsorption, and involves a process where the adsorbate creates a monolayer on the surface of the adsorbent until equilibrium is achieved between the liquid and solid [31]; it also considers that the surface of the adsorbent is homogeneous which has a certain number of available sites which have equal energy capable of adsorbing a single species.
q e = q m a x K L C e 1 + K L C e
The Freundlich isotherm assumes that the adsorption process takes place in multilayers from the exposed surface into the pores of the adsorbent [32].
q e = K F C e 1 / 2
Dubinin–Radushkevich’s model assumes heterogeneous surface and constant adsorption potential in the active centers [33].
q e = q D R × e K D R ε 2
ε = R T × ln 1 + 1 C e
E = 1 2 K D R

2.5. Multicomponent Adsorption

Competitive adsorption tests were performed by placing in contact 2 g/L of adsorbent dose with 50 mL of an equimolar pollutant solution, adding 50 mL of nitrate, sulphate, and phosphate solution at 100 mg/L, for 24 h at the best temperature condition found. The solutions were filtered with 0.45 nm microfilters and syringes and stored in plastic vials and phosphates in amber glass vials, then refrigerated until analysis.

3. Results

3.1. Adsorption Kinetics

To analyze the effect of time on the adsorption process, and to determine the service time of the adsorbent, the adsorption kinetics of the nutrients with the prepared adsorbents were evaluated. Elovich’s pseudo-first and pseudo-second-order models were used to adjust the experimental data and identify possible mechanisms involved in adsorption [34]. The adsorption kinetics are presented in Figure 1 and the model fitting parameters are summarized in Table 1.
From Figure 1, it can be determined that the maximum adsorption capacity of nitrate and sulphate was obtained with B 1:2, and for phosphate with the B 1:3, reaching the adsorption equilibrium close to 180 min in all cases, presenting a fast removal rate during the first minutes. This fast adsorption in the initial stages of the process is due to the influence of the diffusive phenomena from the solution towards the surface of the adsorbent and later to the interior of the pores [35]. Accordingly, with Balarak et al. [36], the high initial concentration of the pollutant at the onset of adsorption helps in creating a better mass transfer driving force, overcoming the resistance to external diffusion; hence, the accelerated adsorption rate.
The adsorption of nitrate and phosphate with the two bio-adsorbents presented a good fitting to the pseudo-first-order model which establishes that the rate of adsorption depends on a mechanism operating at an active site on the surface of the biomass [34]. It was found that the adsorption kinetics of nitrate and phosphate with all biochar follow the pseudo-second-order model, as well as the removal of sulphate with B 1:2, thus establishing that the rate-limiting step is chemisorption due to physicochemical interactions between the two phases [37]. The adjustment of sulphate kinetics for B 1:2 is also described by the Elovich model, thus it is assumed that the limiting step of the process is chemical adsorption [38].

3.2. Adsorption Equilibrium

Adsorption equilibrium was performed by studying the isotherms at 25 °C, initial concentrations of 20, 40, 60, 80, and 100 mg/L. Table 2 summarizes the parameters of adjustment to Langmuir, Freundlich, and Dubinin–Radushkevich’s models by nonlinear regression of the nitrate (N), sulphate (S), and phosphate (P) adsorption isotherms on the synthesized biochar. The best-fitting was chosen according to the correlation coefficient determined (R2).
From Table 2 it is determined that the Freundlich model best adjusts the adsorption data of nitrate and phosphate using B 1:2 and B 1:3; this suggests that the adsorption of nitrate and phosphate is not limited by the monolayer coverage, meaning that multilayer adsorption occurs due to the heterogeneous surface [34]. Sulphate removal data using both adsorbents were described by Dubinin–Radushkevich’s model. The values recorded for E suggest that the process is mostly controlled by the ion exchange mechanism with strong interactions between the active centers and the anion; similarly, the calculated maximum adsorption capacity (qDR) values are closest to the experimental ones. The fitting to Dubinin–Radushkevich’s model assumes that B 1:3 has a heterogeneous structure [39]. Halajnia et al. [40] reported that nitrate and sulphate adsorption was adjusted to Freundlich’s model and phosphate adsorption to Dubinin–Radushkevich’s model. On the other hand, Matusik [41] found that Freundlich’s model described the adsorption of phosphate, nitrate, and sulphate, and the calculated energies of adsorption (E) suggested that ion exchange was dominant for all anions, as in the present study, being all above 8 KJ/mol.
Table 3 summarizes the qmax values obtained in previous studies and the present research for the adsorption of nitrate (NO3), sulfate (SO42−), and phosphate (PO43−) onto adsorbents of different nature. It is observed that the results obtained in the present study for the removal of NO3 are in the interval reported from 5 mg/g up to 68.96 mg/g; for PO43 they are between 19.24 mg/g and 46.67 mg/g, having the adsorbents synthetized in the present study a great performance; finally, the capacities obtained when removing SO42− are in the range of 8.38 and 78.10 mg/g. It is evidenced that the most effective modifications involve the quaternization of bioadsorbents with, epichlorohydrin, pyridine, and quaternary salts as N,N-dimethylformamide, trimethylamine, and diethylamine.

3.3. Multicomponent Adsorption Test Parameters

An experimental set-up was performed in a batch system using a ternary equimolar solution of phosphate, nitrate, and sulphate at 25 °C using 2 g/L g of adsorbent dose in 150 mL of solution (Figure 2).
Figure 2 shows a higher removal capacity of phosphate and sulphate by B 1:3, while nitrate was the least removed material from the solution with similar behavior in the two adsorbents. This behavior of nitrate in the presence of the other two anions may be due to its monovalent nature, and it would have a higher selectivity per specific active center of the bioadsorbent, thus decreasing its removal by competing for the adsorption sites with phosphate and sulphate [54]. It have been demonstrated that adsorbents have higher affinity for higher charge density and divalent anions than for monovalent anions [55]. The selectivity towards phosphate and sulphate over nitrate may be due to the fact that phosphate and sulphate adsorption included electrostatic interactions, surface complexation, and/or ligand exchange by direct bond with hydroxyl groups, as well as inner-sphere complexation as the main binding mechanisms (1). In addition, sulphate had higher adsorption capacity than phosphate and nitrate, which indicated that the number of specific active sites for sulphate ion adsorption are higher than nitrate and phosphate; this may be due to the symmetry (tetrahedral) of the sulphate, as the preferential sorption of sulfate can take place due to its higher charge density, which agrees with what was reported by Vinicius et al., (2). Meanwhile, nitrate adsorption could be attributed to electrostatic interactions rather than surface complexation because it is difficult for nitrate to bond to hydroxyl groups. This behavior has been previously informed by Wu et al. (3), for nitrate and phosphate specifically. The understanding of the reaction mechanisms of the co-adsorption of nitrate, sulphate and phosphate is a very important basis for further optimizing adsorbents as well as the scalability of the process to batch to fixed-bed system; thus, more studies and characterization are needed.

4. Conclusions

This study concluded that: (i) the equilibrium is reached at 180 min. (ii) the maximum adsorption capacity of nitrate and sulphate was obtained with B 1:2 and for phosphate with B 1:3. (iii) The adsorption of nitrate and phosphate with two bioadsorbents, showed a good fitting to the pseudo-first-order and pseudo-second-order model, while that of sulphate on B 1:2 is described by Elovich’s model, establishing that the rate-limiting step is chemisorption due to physicochemical interactions between the two phases. (iv) Freundlich’s model describes the equilibrium of adsorption of the three anions on B 1:1, of nitrate and phosphate using C 1:2 and C 1:3, while Dubinin-Radushkevich adjusts the removal of sulphate on C 1:2 and C 1:3; therefore it is suggested that adsorption occurs in multilayers. (v) The multicomponent study evidences the preference of biochar for phosphate, without indicating competition for the active centers of the material among the anions studied.

Author Contributions

Conceptualization, C.T.-T., A.V.-O. and R.O.-T.; methodology, C.T.-T.; software, A.V.-O.; validation, C.T.-T., A.V.-O. and R.O.-T.; formal analysis, A.V.-O.; investigation, C.T.-T.; resources, C.T.-T.; data curation, R.O.-T.; writing—original draft preparation, A.V.-O.; writing—review and editing, C.T.-T.; visualization, A.V.-O.; supervision, C.T.-T.; project administration, A.V.-O. and R.O.-T.; funding acquisition, C.T.-T. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data that support the findings of this study are available on request from the corresponding author, [C.T.-T.] upon reasonable request.

Acknowledgments

The authors thank to the Universidad de Cartagena to provide the materials, equipment, and laboratory facilities required to successfully conclude this research.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Li, J.; Li, B.; Huang, H.; Zhao, N.; Zhang, M.; Cao, L. Investigation into lanthanum-coated biochar obtained from urban dewatered sewage sludge for enhanced phosphate adsorption. Sci. Total Environ. 2020, 714, 136839. [Google Scholar] [CrossRef] [PubMed]
  2. Yin, Q.; Zhang, B.; Wang, R.; Zhao, Z. Biochar as an adsorbent for inorganic nitrogen and phosphorus removal from water: A review. Environ. Sci. Pollut. Res. 2017, 24, 26297–26309. [Google Scholar] [CrossRef]
  3. Lazaratou, C.; Vayenas, D.; Papoulis, D. The role of clays, clay minerals and clay-based materials for nitrate removal from water systems: A review. Appl. Clay Sci. 2020, 185, 105377. [Google Scholar] [CrossRef]
  4. Ren, Z.; Xu, X.; Wang, X.; Gao, B.; Yue, Q.; Song, W.; Zhang, L.; Wang, H. FTIR, Raman, and XPS analysis during phosphate, nitrate and Cr(VI) removal by amine cross-linking biosorbent. J. Colloid Interface Sci. 2016, 468, 313–323. [Google Scholar] [CrossRef] [PubMed]
  5. Naghizadeh, A.; Ghasemi, F.; Derakhshani, E.; Shahabi, H. Thermodynamic, kinetic and isotherm studies of sulfate removal from aqueous solutions by graphene and graphite nanoparticles. Desalin. Water Treat. 2017, 80, 247–254. [Google Scholar] [CrossRef]
  6. US-EPA; U.S.E.P.A. 2018. Edition of the Drinking Water Standards and Health Advisories Tables. 2018. Available online: https://www.epa.gov/system/files/documents/2022-01/dwtable2018.pdf (accessed on 7 September 2022).
  7. WHO. Nitrate and Nitrite in Drinking-Water. 2019. Available online: https://apps.who.int/iris/bitstream/handle/10665/75380/WHO_SDE_WSH_04.03_56_eng.pdf (accessed on 7 September 2022).
  8. Belkada, F.D.; Kitous, O.; Drouiche, N.; Aoudj, S.; Bouchelaghem, O.; Abdi, N.; Grib, H.; Mameri, N. Electrodialysis for fluoride and nitrate removal from synthesized photovoltaic industry wastewater. Sep. Purif. Technol. 2018, 204, 108–115. [Google Scholar] [CrossRef]
  9. Liu, T.; Hu, S.; Yuan, Z.; Guo, J. High-level nitrogen removal by simultaneous partial nitritation, anammox and nitrite/nitrate-dependent anaerobic methane oxidation. Water Res. 2019, 166, 115057. [Google Scholar] [CrossRef]
  10. Qin, P.; Lu, S.; Liu, X.; Wang, G.; Zhang, Y.; Li, D.; Wan, Z. Removal of tri-(2-chloroisopropyl) phosphate (TCPP) by three types of constructed wetlands. Sci. Total Environ. 2020, 749, 141668. [Google Scholar] [CrossRef]
  11. Nur, H.M.; Yüzer, B.; Aydin, M.I.; Aydin, S.; Öngen, A.; Selçuk, H. Desalination and fate of nutrient transport in domestic wastewater using electrodialysis membrane process. Desalin. Water Treat. 2019, 172, 323–329. [Google Scholar] [CrossRef]
  12. Liu, R.; Sui, Y.; Wang, X. Metal–organic framework-based ultrafiltration membrane separation with capacitive-type for enhanced phosphate removal. Chem. Eng. J. 2019, 371, 903–913. [Google Scholar] [CrossRef]
  13. Boeykens, S.P.; Piol, M.N.; Legal, L.S.; Saralegui, A.B.; Vázquez, C. Eutrophication decrease: Phosphate adsorption processes in presence of nitrates. J. Environ. Manag. 2017, 203, 888–895. [Google Scholar] [CrossRef] [PubMed]
  14. Wu, K.; Li, Y.; Liu, T.; Huang, Q.; Yang, S.; Wang, W.; Jin, P. The simultaneous adsorption of nitrate and phosphate by an organic-modified aluminum-manganese bimetal oxide: Adsorption properties and mechanisms. Appl. Surf. Sci. 2019, 478, 539–551. [Google Scholar] [CrossRef]
  15. Cao, W.; Dang, Z.; Zhou, X.-Q.; Yi, X.-Y.; Wu, P.; Zhu, N.-W.; Lu, G.-N. Removal of sulphate from aqueous solution using modified rice straw: Preparation, characterization and adsorption performance. Carbohydr. Polym. 2011, 85, 571–577. [Google Scholar] [CrossRef]
  16. Alagha, O.; Manzar, M.S.; Zubair, M.; Anil, I.; Mu’Azu, N.D.; Qureshi, A. Comparative Adsorptive Removal of Phosphate and Nitrate from Wastewater Using Biochar-MgAl LDH Nanocomposites: Coexisting Anions Effect and Mechanistic Studies. Nanomaterials 2020, 10, 336. [Google Scholar] [CrossRef]
  17. Li, J.-H.; Lv, G.-H.; Bai, W.-B.; Liu, Q.; Zhang, Y.-C.; Song, J.-Q. Modification and use of biochar from wheat straw (Triticum aestivumL.) for nitrate and phosphate removal from water. Desalin. Water Treat. 2016, 57, 4681–4693. [Google Scholar] [CrossRef]
  18. Zhao, S.; Wang, B.; Gao, Q.; Gao, Y.; Liu, S. Adsorption of phosphorus by different biochars. Spectrosc. Lett. 2017, 50, 73–80. [Google Scholar] [CrossRef]
  19. Shukla, N.; Sahoo, D.; Remya, N. Biochar from microwave pyrolysis of rice husk for tertiary wastewater treatment and soil nourishment. J. Clean. Prod. 2019, 235, 1073–1079. [Google Scholar] [CrossRef]
  20. Wang, Z.; Shen, D.; Shen, F.; Li, T. Phosphate adsorption on lanthanum loaded biochar. Chemosphere 2016, 150, 1–7. [Google Scholar] [CrossRef]
  21. Zare, L.; Ghasemi-Fasaei, R. Investigation of equilibrium isotherm and kinetic modeling to asses sorption characteristics of nitrate onto palm leaf biochar. Iran. J. Chem. Chem. Eng. 2018, 38, 143–153. [Google Scholar]
  22. Manjunath, S.V.; Kumar, M. Evaluation of single-component and multi-component adsorption of metronidazole, phosphate and nitrate on activated carbon from Prosopıs julıflora. Chem. Eng. J. 2018, 346, 525–534. [Google Scholar] [CrossRef]
  23. Villabona-Ortiz, Á.; Tejada-Tovar, C.; Ortega-Toro, R. Effect of temperature in removing of anions in solution on biochar using Zea mays stalks as a precursor. J. Water L. Dev. 2021, 50, 64–68. [Google Scholar] [CrossRef]
  24. ASTM. ASTM D 4130-15; Standard Test Method for Sulfate in Brackish Water, Seawater, and Brines. ASTM: West Conshohocken, PA, USA, 2018; pp. 1–5. [Google Scholar]
  25. ASTM. ASTM D7781-14; Standard Test Method for Nitrite-Nitrate in Water by Nitrate Reductase. ASTM: West Conshohocken, PA, USA, 2018; pp. 1–8. [Google Scholar]
  26. ASTM. ASTM D 515-60; Standard Test Method for Phosphate Ion in Water. ASTM: West Conshohocken, PA, USA, 2018; pp. 1–4. [Google Scholar]
  27. Lonappan, L.; Rouissi, T.; Brar, S.K.; Verma, M.; Surampalli, R.Y. An insight into the adsorption of diclofenac on different biochars: Mechanisms, surface chemistry, and thermodynamics. Bioresour. Technol. 2018, 249, 386–394. [Google Scholar] [CrossRef] [PubMed]
  28. Hubbe, M.A.; Azizian, S.; Douven, S. Implications of apparent pseudo-second-order adsorption kinetics onto cellulosic materials: A review. BioResources 2019, 14, 7582–7626. [Google Scholar] [CrossRef]
  29. Riahi, K.; Chaabane, S.; Ben Thayer, B. A kinetic modeling study of phosphate adsorption onto Phoenix dactylifera L. date palm fibers in batch mode. J. Saudi Chem. Soc. 2017, 21, S143–S152. [Google Scholar] [CrossRef]
  30. Naushad, M.; Sharma, G.; Kumar, A.; Sharma, S.; Ghfar, A.A.; Bhatnagar, A.; Stadler, F.J.; Khan, M.R. Efficient removal of toxic phosphate anions from aqueous environment using pectin based quaternary amino anion exchanger. Int. J. Biol. Macromol. 2018, 106, 1–10. [Google Scholar] [CrossRef] [PubMed]
  31. Rasmey, A.-H.M.; Aboseidah, A.A.; Youssef, A.K. Application of langmuir and freundlich isotherm models on biosorption of Pb2+ by freez-dried biomass of Pseudomonas aeruginosa. Egypt. J. Microbiol. 2018, 53, 37–48. [Google Scholar] [CrossRef]
  32. Xi, Y.; Huang, M.; Luo, X. Enhanced phosphate adsorption performance by innovative anion imprinted polymers with dual interaction. Appl. Surf. Sci. 2019, 467, 135–142. [Google Scholar] [CrossRef]
  33. Rodiguez, M.H.; Yperman, J.; Carleer, R.; Maggen, J.; Dadi, D.; Gryglewicz, G.; Van der Bruggen, B.; Hernández, J.F.; Otero-Calvis, A. Adsorption of Ni(II) on spent coffee and coffee husk based activated carbon. J. Environ. Chem. Eng. 2018, 6, 1161–1170. [Google Scholar] [CrossRef]
  34. Wang, L.; Xu, Z.; Fu, Y.; Chen, Y.; Pan, Z.; Wang, R.; Tan, Z. Comparative analysis on adsorption properties and mechanisms of nitrate and phosphate by modified corn stalks. RSC Adv. 2018, 8, 36468–36476. [Google Scholar] [CrossRef]
  35. Bashir, M.T.; Ali, S.; Idris, A.; Haroon, R. Kinetic and thermodynamic study of nitrate adsorption from aqueous solution by lignocellulose-based anion resins. Desalination Water Treat. 2017, 62, 449–456. [Google Scholar] [CrossRef]
  36. Balarak, D.; Zafariyan, M.; Igwegbe, C.A.; Onyechi, K.K.; Ighalo, J.O. Adsorption of acid blue 92 dye from aqueous solutions by single-walled carbon nanotubes: Isothermal, kinetic, and thermodynamic studies. Environ. Process. 2021, 8, 869–888. [Google Scholar] [CrossRef]
  37. Golie, W.M.; Upadhyayula, S. An investigation on biosorption of nitrate from water by chitosan based organic-inorganic hybrid biocomposites. Int. J. Biol. Macromol. 2017, 97, 489–502. [Google Scholar] [CrossRef]
  38. William Kajjumba, G.; Emik, S.; Öngen, A.; Kurtulus Özcan, H.; Aydın, S. Modelling of adsorption kinetic processes—Errors, theory and application. In Advanced Sorption Process Applications; Intechopen: London, UK, 2019; pp. 1–19. [Google Scholar] [CrossRef]
  39. Karthikeyan, P.; Meenakshi, S. Synthesis and characterization of Zn–Al LDHs/activated carbon composite and its adsorption properties for phosphate and nitrate ions in aqueous medium. J. Mol. Liq. 2019, 296, 111766. [Google Scholar] [CrossRef]
  40. Halajnia, A.; Oustan, S.; Najafi, N.; Khataee, A.; Lakzian, A. Adsorption–desorption characteristics of nitrate, phosphate and sulfate on Mg–Al layered double hydroxide. Appl. Clay Sci. 2013, 80, 305–312. [Google Scholar] [CrossRef]
  41. Matusik, J. Arsenate, orthophosphate, sulfate, and nitrate sorption equilibria and kinetics for halloysite and kaolinites with an induced positive charge. Chem. Eng. J. 2014, 246, 244–253. [Google Scholar] [CrossRef]
  42. Qiao, H.; Mei, L.; Chen, G.; Liu, H.; Peng, C.; Ke, F.; Hou, R.; Wan, X.; Cai, H. Adsorption of nitrate and phosphate from aqueous solution using amine cross-linked tea wastes. Appl. Surf. Sci. 2019, 483, 114–122. [Google Scholar] [CrossRef]
  43. Barroso-Solares, S.; Merillas, B.; Cimavilla-Román, P.; Rodriguez-Perez, M.; Pinto, J. Enhanced nitrates-polluted water remediation by polyurethane/sepiolite cellular nanocomposites. J. Clean. Prod. 2020, 254, 120038. [Google Scholar] [CrossRef]
  44. Banu, H.A.T.; Karthikeyan, P.; Meenakshi, S. Comparative studies on revival of nitrate and phosphate ions using quaternized corn husk and jackfruit peel. Bioresour. Technol. Rep. 2019, 8, 100331. [Google Scholar] [CrossRef]
  45. Xia, F.; Yang, H.; Li, L.; Ren, Y.; Shi, D.; Chai, H.; Ai, H.; He, Q.; Gu, L. Enhanced nitrate adsorption by using cetyltrimethylammonium chloride pre-loaded activated carbon. Environ. Technol. 2019, 41, 3562–3572. [Google Scholar] [CrossRef]
  46. Zhao, B.; Xu, H.; Ma, F.; Zhang, T.; Nan, X. Effects of dairy manure biochar on adsorption of sulfate onto light sierozem and its mechanisms. RSC Adv. 2019, 9, 5218–5223. [Google Scholar] [CrossRef] [Green Version]
  47. Qiu, L.; Burton, G.R.; Rousseau, S.; Qian, J. Kinetics and thermodynamics of sulfate adsorption on magnetite at elevated temperatures. J. Solut. Chem. 2019, 48, 1488–1502. [Google Scholar] [CrossRef]
  48. Salami, A.H.; Bonakdari, H.; Akhbari, A.; Shamshiri, A.; Mousavi, S.F.; Farzin, S.; Hassanvand, M.R.; Noori, A. Performance assessment of modified clinoptilolite and magnetic nanotubes on sulfate removal and potential application in natural river samples. J. Incl. Phenom. Macrocycl. Chem. 2020, 97, 51–63. [Google Scholar] [CrossRef]
  49. Ao, H.; Cao, W.; Hong, Y.; Wu, J.; Wei, L. Adsorption of sulfate ion from water by zirconium oxide-modified biochar derived from pomelo peel. Sci. Total Environ. 2020, 708, 135092. [Google Scholar] [CrossRef]
  50. Runtti, H.; Tynjälä, P.; Tuomikoski, S.; Kangas, T.; Hu, T.; Rämö, J.; Lassi, U. Utilisation of barium-modified analcime in sulphate removal: Isotherms, kinetics and thermodynamics studies. J. Water Process. Eng. 2017, 16, 319–328. [Google Scholar] [CrossRef]
  51. Sereshti, H.; Afsharian, E.Z.; Bidhendi, M.E.; Nodeh, H.R.; Kamboh, M.A.; Yilmaz, M. Removal of phosphate and nitrate ions aqueous using strontium magnetic graphene oxide nanocomposite: Isotherms, kinetics, and thermodynamics studies. Environ. Prog. Sustain. Energy 2020, 39, e13332. [Google Scholar] [CrossRef]
  52. Dong, S.; Ji, Q.; Wang, Y.; Liu, H.; Qu, J. Enhanced phosphate removal using zirconium hydroxide encapsulated in quaternized cellulose. J. Environ. Sci. 2020, 89, 102–112. [Google Scholar] [CrossRef] [PubMed]
  53. Huang, Y.; Lee, X.; Grattieri, M.; Yuan, M.; Cai, R.; Macazo, F.C.; Minteer, S.D. Modified biochar for phosphate adsorption in environmentally relevant conditions. Chem. Eng. J. 2020, 380, 122375. [Google Scholar] [CrossRef]
  54. Iftekhar, S.; Küçük, M.E.; Srivastava, V.; Repo, E.; Sillanpää, M. Application of zinc-aluminium layered double hydroxides for adsorptive removal of phosphate and sulfate: Equilibrium, kinetic and thermodynamic. Chemosphere 2018, 209, 470–479. [Google Scholar] [CrossRef]
  55. Constantino, L.V.; Quirino, J.N.; Monteiro, A.M.; Abrão, T.; Parreira, P.S.; Urbano, A.; Santos, M.J. Sorption-desorption of selenite and selenate on Mg-Al layered double hydroxide in competition with nitrate, sulfate and phosphate. Chemosphere 2017, 181, 627–634. [Google Scholar] [CrossRef]
Figure 1. Adsorption kinetics of (a) nitrate, (b) sulphate and (c) phosphate; source: own study.
Figure 1. Adsorption kinetics of (a) nitrate, (b) sulphate and (c) phosphate; source: own study.
Water 14 02906 g001aWater 14 02906 g001b
Figure 2. Competitive adsorption of nitrate, sulphate, and phosphate; source: own study.
Figure 2. Competitive adsorption of nitrate, sulphate, and phosphate; source: own study.
Water 14 02906 g002
Table 1. Adsorption kinetic parameters of nitrate, sulphate, and phosphate.
Table 1. Adsorption kinetic parameters of nitrate, sulphate, and phosphate.
ModelParameterNitrateSulphatePhosphate
B 1:2B 1:3B 1:2B 1:3B 1:2B 1:3
Pseudo-first orderqe10.649.098.6211.0823.5918.12
k10.030.020.020.000390.010.008
R20.960.960.870.790.990.97
Pseudo-second orderk211.5310.059.9010.8326.8320.95
qe0.0040.0030.0018.78 × 10−45.26 × 10−44.82 × 10−4
R20.930.940.940.870.990.97
Elovichβ0.530.550.590.550.210.29
α1.330.640.530.381.080.75
R20.810.880.970.940.940.93
Table 2. Modelling adjustment parameters of the Cr (VI) adsorption isotherm.
Table 2. Modelling adjustment parameters of the Cr (VI) adsorption isotherm.
Isotherm ModelParametersNitrateSulphatePhosphate
B 1:2B 1:3B 1:2B 1:3B 1:2B 1:3
Langmuirqmax (mg/g)9788.17487.9328.687625.078400.4973834.8
KL (L/mg)3.68 × 10−52.73 × 10−40.012.28 × 10−53.77 × 10−62.99 × 10−6
R20.990.950.760.890.590.61
Freundlichkf (mg/g (L/mg)1/n)0.130.120.450.117.211.4 × 10−4
n0.980.981.330.900.390.35
R20.990.950.710.900.850.89
Dubinin–RadushkevichqDR (mg/g)11.9312.0312.6415.0344.55847.75
KDR (mol2/kJ2)1.46 × 10−41.71 × 10−49.02 × 10−51.65 × 10−43.13 × 10−45.41
E (KJ/mol)58.5454.0174.4555.0139.9930.39
R20.820.890.860.990.790.79
Table 3. Modelling adjustment parameters of the Cr (VI) adsorption isotherm.
Table 3. Modelling adjustment parameters of the Cr (VI) adsorption isotherm.
PollutantAdsorbentqmax (mg/g)Reference
NitrateZn–Al LDHs/activated carbon composite73.742[39]
Amine crosslinked tea waste136.43[42]
Polyurethane/sepiolite cellular nanocomposites23.30[43]
Corn husk quaternized with N,N-dimethylformamide, ethylenediamine, and triethylamine79.09[44]
Jackfruit peel quaternized with N,N-dimethylformamide, ethylenediamine, and triethylamine62.91
Cetylpyridinium bromide modified zeolite28.06[16]
Unmodified activated carbon3.86[45]
25% CTAC-modified activated carbon7.1
Activated carbon modified with CTAC at 50%.10.5
100% CTAC-modified activated carbon14.3
SulphateManure biochar78.4[46]
Mangenite16.4[47]
Clinoptilolite74.63[48]
Magnetic nanotubes94
Zirconium oxide-modified biochar derived from pomelo peel35.21[49]
Pomelo peel1.02
Barium-modified analcime2.3[50]
Barium-modified acid-washed analcime13.7
Barium-modified zeolite3.8
PhosphateStrontium magnetic graphene oxide nanocomposite238.09[51]
Zirconium hydroxide encapsulated in quaternized cellulose83.6[52]
Lanthanum-coated biochar from urban dewatered sewage sludge93.91[1]
Biochar from wheat straw modified with chitosan77.61[53]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Villabona-Ortíz, A.; Ortega-Toro, R.; Tejada-Tovar, C. Selective and Competitive Adsorption of Anions in Solution on Porous Adsorbent from Zea mays Steams: Kinetic and Equilibrium Study. Water 2022, 14, 2906. https://doi.org/10.3390/w14182906

AMA Style

Villabona-Ortíz A, Ortega-Toro R, Tejada-Tovar C. Selective and Competitive Adsorption of Anions in Solution on Porous Adsorbent from Zea mays Steams: Kinetic and Equilibrium Study. Water. 2022; 14(18):2906. https://doi.org/10.3390/w14182906

Chicago/Turabian Style

Villabona-Ortíz, Angel, Rodrigo Ortega-Toro, and Candelaria Tejada-Tovar. 2022. "Selective and Competitive Adsorption of Anions in Solution on Porous Adsorbent from Zea mays Steams: Kinetic and Equilibrium Study" Water 14, no. 18: 2906. https://doi.org/10.3390/w14182906

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