Next Article in Journal
On the Retrieval of the Water Quality Parameters from Sentinel-3/2 and Landsat-8 OLI in the Nile Delta’s Coastal and Inland Waters
Next Article in Special Issue
Influence of Sulfate Reduction on Arsenic Migration and Transformation in Groundwater Environment
Previous Article in Journal
Systematic Application of Sponge City Facilities at Community Scale Based on SWMM
Previous Article in Special Issue
A Comparative Study of the Growth and Nutrient Removal Effects of Five Green Microalgae in Simulated Domestic Sewage
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Fluoride Adsorption Comparison from Aqueous Solutions Using Al- and La-Modified Adsorbent Prepared from Polygonum orientale Linn.

1
Department of Biological Science and Technology, Jinzhong University, Jinzhong 030600, China
2
Shanxi Key Laboratory for Research and Development of Regional Plants, School of Life Science, Shanxi University, Taiyuan 030006, China
*
Author to whom correspondence should be addressed.
Water 2022, 14(4), 592; https://doi.org/10.3390/w14040592
Submission received: 21 January 2022 / Revised: 11 February 2022 / Accepted: 14 February 2022 / Published: 15 February 2022
(This article belongs to the Special Issue Water Pollution and Its Impact on Human Health)

Abstract

:
Al- and La-modified adsorbent materials (PO–Al, PO–La) were prepared by impregnating Polygonum orientale Linn. straw with Al2(SO4)3 and La(NO3)3·6H2O solutions. The potential of removing fluoride using these modified adsorbents was examined. In the PO, PO–Al and PO–La adsorption systems, the fluoride adsorption process followed pseudo-second-order kinetics, and the kinetic constants for k2 and R2 were 0.0276 and 0.9609; 0.2070 and 0.9994; 0.1266 and 0.9933, respectively. The adsorption equilibrium results showed the best match with Langmuir isotherms. Moreover, the maximum monolayer adsorption capacity of PO, PO–Al and PO–La are 0.0923, 3.3190 and 1.2514 mg/g, respectively, in 30 °C. The regeneration results show that the effectively regenerating ability of modified adsorbents. Al-modified adsorbent showed the best results in terms of cost-effectiveness and adsorption efficiency for fluoride adsorption.

1. Introduction

As stated by the report of the World Health Organization, the intake of fluoride in drinking water at low concentrations (0.5–1 mg/L) is beneficial, especially in promoting enamel calcification and protecting teeth against teeth decay [1,2,3,4]. However, when long-time exposure exceeds 1.5 mg/L fluoride concentrations, it can lead to fluorosis, with a range of serious consequences, including impaired kidney function, bone cancer, reduced immunity, digestive and nervous disorders, endocrine gland lesions, and other conditions [5,6,7,8]. Fluorosis caused by high concentrations of fluorine has been reported almost everywhere except in East Asia and Australia [5,9]. Accordingly, excessive fluoride contamination in groundwater has been recognized as an extremely serious problem worldwide.
Various techniques have been developed to remove excess fluoride from drinking water, such as adsorption, coagulation, ion exchange, microbial-induced precipitation [10,11], electrocoagulation [12], electrostatic unit [13] and electrodialysis [5,14]. Among these techniques, adsorption is effective, economical, convenient, and environmentally friendly [3,4,15,16]. Biosorption is a new technology for water treatment using a large amount of available biological materials. The biosorbents for fluoride removal contain chitin, chitosan, etc. The regenerated bone char media was optimized and applied to the defluoridation of drinking water in Tanzania, with the highest fluoride removal and adsorption capacity [17]. Therefore, there is a need to research and develop novel technologies for fluoride removal that use low-cost biomass containing carboxylic acid and amine functional groups that could be applied to water pollution control [18,19,20].
Polygonum orientale Linn., belonging to Polygonum genus, Polygonaceae family, is a fast-growing annual herbage, commonly found in lake edges, drainage ditches, marshes, and other wetlands [21]. It could adsorb large quantities of nutrients and could produce large amounts of biomass at a low cost [22,23,24]. Therefore, this plant provides a good basis for the study of modified materials as an adsorbent.
The surface adsorption performance of modified adsorbents could be enhanced by changing their surface chemistry and pore structure [4,25]. Many modification methods have been applied to improve affinities towards contaminants in various types of water [26,27], furnishing satisfactory results, particularly using surface-modified adsorbents and/or mineral-based [3].
(1) Modification of alumina with Al3+. Over the years, activated alumina has been proven by researchers to effectively remove fluoride, and a series of explorations have been carried out on it. For example, in order to explore the effects of different experimental conditions on the removal ability of fluoride in water during the experiment, Ku and Chiou [28] studied the conditions of temperature, pH, and initial concentration during the experiment. At the same time, some studies have also explored the interaction process and mechanism between F and amorphous Al(OH)3, Al(OH)3, and Al2O3 [29]. Meanwhile, the researchers also modified the surface of Al2O3 to improve the adsorption effect of activated Al2O3, such as adding La3+ and Y3+ to Al2O3 to increase its surface active adsorption sites [30]; using MnO2 modification is used to prepare Mn–Al2O3 oxide coating (MOCA) for fluorine removal, and Al2O3-supported La(OH)3 is modified to remove fluoride in water [31,32]. The results have shown that after modification of Al2O3 impregnated with La(OH)3, the adsorption capacity for F is the highest, the maximum fluoride uptake capacity for MOCA was found to be 2.85 mg/g [31]. This may be because the ion exchange between fluoride and OH on the surface of Al2O3 promotes the F removal process.
(2) Modification of alumina with La3+. The biomass materials are widely used as adsorbents because of their large reserves, low cost, and good adsorption capacity. So, some researchers reported that the use of lanthanide compounds to modify the surface of chitosan can improve the ability to remove F [33].
In addition, Kamble et al. [32] and Parmar et al. [34] studied the difference in the adsorption of fluoride with La(OH)3 and AlCl3 modification; the results showed that after La(OH)3 and AlCl3 modification, the treatment had better adsorption ability of F. Similar research also includes the use of agricultural biomass materials such as coconut husk and rice husk to remove fluoride [35]. Therefore, this study aimed to assess the feasibility and adsorption capacity of Al2(SO4)3- and La(NO3)3·6H2O-modified biomaterials derived from P. orientale (PO, PO–Al, PO–La) to remove fluoride in aqueous solutions. Furthermore, the ability of these adsorbents to perform desorption test was examined for potential industrial application.

2. Materials and Methods

2.1. Preparation of Modified Adsorbents

Polygonum orientale Linn. was gathered from Fenhe river (Shanxi province, northern China), after the process of cutting, washing, and drying in an oven (101A–3, Boxun, Shanghai, China), the samples were ground into powder by a grinder and then sieved through a standard sieve to obtain 160–200 mesh particles. This precursor (2 g) was added to Al2(SO4)3 (3%) or La(NO3)3·6H2O (5%) aqueous solutions (200 mL). After soaking for 24 h, the solution was filtered by vacuum filtration and the resultant modified adsorbents of Al2(SO4)3-modified (PO–Al) and La(NO3)3·6H2O-modified (PO–La) were obtained after washed with distilled water until the pH was 7.0 and dried at 120 °C for 8 h.

2.2. Adsorption Experiments

Fluoride stock solution was prepared by dissolving NaF (2.21 g) in 1000 mL deionized water. The solution pH was adjusted by adding 0.5 M NaOH or 0.5 M HNO3 solutions.
Adsorption experiments were performed by shaking prepared modified adsorbent (2.0 g) with NaF experimental solutions (100 mL) of different concentrations (10, 20, and 30 mg/L). Continuous stirring was performed throughout the experiment at a constant speed of 120 rpm. In addition, the adsorption isotherm tests were performed at 10, 20, and 30 °C, simultaneously, other things being equal. Then, the supernatant liquids were filtered by a disposable injector and syringe filter (13 mm × 0.45 µm, Millipore, Billerica, MA, USA), and the F concentration was measured by F concentration meter (PXSJ-216F, Leici, Shanghai, China), the F concentration in each flask was determined three times.
In this study, two isotherm models of Langmuir and Freundlich were utilized to analyze the experimental equilibrium data for fluoride adsorption by PO, PO–Al, and PO–La. The adsorption isotherm model is a basic model describing the interactions between adsorbate and adsorbent.

2.2.1. Langmuir Isotherm

The Langmuir isotherm model formula as shown in the following Equation (1) [36]:
C e q e = 1 Q m b + 1 Q m C e
where Ce (mg/L) represents the equilibrium concentration of the adsorbate, qe (mg/g) is the maximum adsorption capacity on the equilibrium phase, b (L/mg) is the Langmuir adsorption constant, and Qm (mg/g) is the maximum amount adsorbed of adsorbent. However, a dimensionless equilibrium parameter RL could be defined according to Equation (7) to determine whether adsorption is beneficial [37]:
R L = 1 1 + b C 0
where RL indicates whether the isotherm is favorable (0 < RL < 1), irreversible (RL = 0), linear (RL = 1), or unfavorable (RL > 1). b (L/mg) is the Langmuir isotherm constant, and C0 (mg/L) is the initial fluoride concentration.

2.2.2. Freundlich Isotherm

The Freundlich isotherm model could be expressed in Equation (3) [38]:
log q e = log K F + ( 1 n ) log C e
where KF ((mg/g) (L/mg)1/n) indicates the adsorption capacity of the adsorbent, with a larger KF value indicating a larger adsorption capacity. The Freundlich isotherm constant n indicates the favorable degree of the adsorption process, with 1/n < 1 indicating the normal Freundlich isotherm and cooperative adsorption process.

2.3. Adsorption Kinetic Models

2.3.1. Pseudo-First-Order Model

The pseudo-first-order kinetic model is shown in Equation (4) [39].
d q t d t = k 1 ( q e q t )
When qt = 0 at t = 0, Equation (4) could be integrated to give Equation (5):
log   ( q e q t ) = log q e k 1 2.303 t
where qe and qt (mg/g) are the adsorption capacity of fluoride at equilibrium and time t, respectively, t (min) is the adsorption time, and k1 (min−1) is a rate constant for this equation.

2.3.2. Pseudo-Second-Order Model

The pseudo-second-order kinetic model as shown in Equation (6) [40]:
d q t d t = k 2 ( q e q t ) 2  
when Equation (6) is simplified and qt = 0 at t = 0, the new equation could be rearranged into Equation (7):
t q t = 1 k 2 q e 2 + 1 q e t  
h = k 2 q e 2  
where h (mg/ (g min)) represents the product of the pseudo-second-order kinetic constant k2 (g/ (mg min)) and the maximum adsorption capacity on the equilibrium phase qe (mg/g), and it represents the initial adsorption rate.

2.4. Desorption Studies

The modified adsorbents (PO, PO–Al, PO–La) was inputted to a 100 mg/L NaF solution and kept at a constant temperature in a shaker at 150 rpm until equilibrium was reached. The concentration Cad of fluoride was defined as the difference between the initial and equilibrium fluoride concentration (Co − Ce). The spent adsorbent powder was washed with deionized water, over dried at 40 °C, 10 h and mixed with H2O, 0.1 M HCl, NaOH, and 95% ethanol to desorb the fluoride. The fluoride concentration Cde (mg/L) in the solution was determined after the desorption process. The desorption capacity of fluoride was calculated using Equation (9):
Desorption   ( % ) = C d e C a d   ×   100

3. Results

3.1. Adsorption Kinetics

By surface modification of the adsorbent, the adsorption capacity and strength of the adsorbent can be improved without a significant increase in cost. These modified adsorbents were prepared by impregnating P. orientale with Al2(SO4)3 and La(NO3)3·6H2O solutions for removing fluoride in aqueous solutions.
In this section, pseudo-first-order and pseudo-second-order kinetic models are used to fit the adsorption process of fluoride onto three types of adsorbents. Meanwhile, kinetic process fitting is very important to elucidate the adsorption kinetic mechanism and diffusion process. The parameters of the pseudo-first-order and pseudo-second models are shown in Table 1. The comparison of different kinetic models for the adsorption of fluorine at different concentrations of PO, PO–Al, and PO–La are shown in Figure 1 and Figure 2. The pseudo-first-order kinetic curves were poorly fitted to the experimental kinetic data. This disagreement was also confirmed by the low R2. Therefore, it was suggested that fluoride adsorption onto the biochar activated carbon did not conform to the pseudo-first-order model. In contrast, the results showed a high degree fitting degree with the pseudo-second-order model. This great agreement was further supported by the similar calculated and experimental values of qe.
These results show that the adsorption mechanism may rely not only on the interaction with fluoride, but also on the role of adsorbent [41], and the rate-limiting step might be chemisorption, involving valence forces by the exchange or sharing of electrons [26,27].

3.2. Adsorption Isotherms

The adsorption isotherm of fluoride adsorption onto PO, PO–Al, and PO–La at 283, 293 and 303 K, respectively, based on the linear forms of the Langmuir and Freundlich isotherms are shown in Figure 3 and Figure 4, respectively. The isotherm constants and correlation coefficients calculated from the linear forms of the two isotherms are summarized in Table 2. The experimental data of fluoride adsorption onto PO and PO–La were fitted to the Langmuir isotherm model. The adsorption of F on PO–Al was fitted well by both the Langmuir and Freundlich isotherm models, while a better fit with the Langmuir equation was statistically confirmed by the larger R2 values closer to unity. Furthermore, the results show that the adsorption process is favorable, because the Langmuir isotherm value of RL was between 0 and 1. In addition, the maximum adsorption capacity of Qm, improved with increasing temperature, revealing an endothermic process. Adsorption capacity showed that PO, PO–Al, and PO–La could be employed as promising adsorbents to remove fluoride from an aqueous solution. In addition, Table 3 lists the comparison of the fluoride adsorption capacity in this study and reported in other literature. These values are based on the Langmuir adsorption capacity of the adsorbent. The results show that in the field of low-cost adsorbents, the adsorbents prepared in this study are superior to most of the adsorbents in removing fluoride.

3.3. Desorption

A suitable defluorine adsorbent should not only have high defluorine removal capacity and cost efficiency, but also be easy to desorb F and be able to be regenerated efficiently so that the adsorbent can be reused many times. The desorbent should not cause any damage to the adsorbent, resulting in a decline in adsorption capacity. For economic and environmental reasons, the only reusable sorbent can be of practical value in practical systems.
Desorption of F is performed by leaching of adsorbed F by different solvents as eluants, which could contribute to elucidate the mechanism of adsorption and assess the stability of the loaded adsorbents. The elution of fluoride from the adsorbent by water, strong base or acid, and organic substance symbolized the adsorption of fluoride onto PO by the weak force, ion exchange, and chemical adsorptions, behaviors, respectively. Figure 5 showed the results of the desorption percentages using 0.1 M NaOH, deionized water, 0.1 M HCl, and 95% ethanol were measured for PO, PO–Al, PO–La, respectively. If the strong acid or base could desorb a large amount of adsorbate from the adsorbent, the adsorbate is attached to the adsorbent by ion exchange. The organic solvent, for instance, ethanol, could elute F from PO–Al easily and indicated that chemical adsorption is the main phenomenon of PO–Al adsorption of F. Strong chemical bonds between the adsorbate molecules and the adsorbent surface may not be broken by the eluting agents. NaOH displayed certain desorption effect on F. This indicated F has a strong chemical attachment to PO–Al, which can be broken by thermal forces.

4. Conclusions

The modified adsorbent showed the most potential as an economic and efficient adsorbent for fluoride adsorption from aqueous solutions. Meanwhile, the pseudo-second-order kinetic and Langmuir model indicated that the adsorption process of fluoride ions was mainly chemical adsorption. The Langmuir model showing that adsorption was a spontaneous and endothermic process. The increasing temperature was beneficial to adsorption. In the reusability test, ethanol and NaOH were effective materials for regenerating the adsorbent. Furthermore, desorbed PO–Al was highly effective in fluoride resorption. Despite these, more research is needed to translate the process to an industrial scale. In addition, regeneration studies need to be carried out to more extent to recover the adsorbent and improve the economic viability of the process.

Author Contributions

Funding acquisition, S.X. and J.F.; Methodology, Q.Z. and A.L.; Visualization, K.Z.; Writing—original draft, S.S. All authors have read and agreed to the published version of the manuscript.

Funding

We thank for the Social Development Foundation of Shanxi (No. 201603D321001) funding this research.

Acknowledgments

We are grateful to Simon Partridge (Imperial College London) for his editorial.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mousny, M.; Omelon, S.; Wise, L.; Everett, E.T.; Dumitriu, M.; Holmyard, D.P.; Banse, X.; Devogelaer, J.P.; Grynpas, M.D. Fluoride effects on bone formation and mineralization are influenced by genetics. Bone 2008, 43, 1067–1074. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Sundaran, C.S.; Viswanathan, N.; Meenakshi, S. Uptake of fluoride by nanohydroxyapatite/chitosan: –bioinorganic composite. Bioresour. Technol. 2008, 99, 8226–8230. [Google Scholar] [CrossRef] [PubMed]
  3. Bhatnagara, A.; Kumara, E.; Sillanpaa, M. Fluoride removal from water by adsorption–A review. Chem. Eng. J. 2011, 171, 811–840. [Google Scholar] [CrossRef]
  4. Loganathan, P.; Vigneswaran, S.; Kandasamy, J.; Naidu, R. Defluoridation of drinking water using adsorption processes. J. Hazard. Mater. 2013, 248–249, 1–19. [Google Scholar] [CrossRef] [PubMed]
  5. Miretzky, P.; Cirelli, A.F. Fluoride removal from water by chitosan derivatives and composites: A review. J. Fluorine Chem. 2011, 132, 231–240. [Google Scholar] [CrossRef]
  6. Islam, M.; Patel, R.K. Evaluation of removal of fluoride from aqueous solution using quick lime. J. Hazard. Mater. 2007, 143, 303–310. [Google Scholar] [CrossRef] [PubMed]
  7. Paudyal, H.; Pangeni, B.; Inoue, K.; Kawakita, H.; Ohto, K.; Alam, S. Adsorptive removal of fluoride from aqueous medium using a fixed bed column packed with Zr (IV) loaded dried orange juice residue. Bioresource Technol. 2013, 146, 713–720. [Google Scholar] [CrossRef]
  8. Paudyal, H.; Pangeni, B.; Inoue, K.; Kawakita, H.; Ohto, K.; Ghimire, K.N.; Alam, S. Preparation of novel alginate–based anion exchanger from Ulva japonica and its application for the removal of trace concentrations of fluoride from water. Bioresource Technol. 2013, 148, 221–227. [Google Scholar] [CrossRef]
  9. Paudyal, H.; Pangeni, B.; Ghimire, K.N.; Inoue, K.; Kawakita, H.; Ohto, K.; Alam, S. Adsorption behavior of orange waste gel for some rare earth ions and its application to the removal of fluoride from water. Chem. Eng. J. 2012, 195, 289–296. [Google Scholar] [CrossRef]
  10. Liu, J.; Su, J.F.; Ali, A.; Wang, Z.; Zhang, R.J. Potential of a novel facultative anaerobic denitrifying Cupriavidus sp. W12 to remove fluoride and calcium through calcium bioprecipitation. J. Hazard. Mater. 2022, 423, 126976. [Google Scholar] [CrossRef]
  11. Wang, Z.; Su, J.F.; Ali, A.; Zhang, R.J.; Yang, W.S.; Xu, L.; Zhao, T.B. Microbially induced calcium precipitation based simultaneous removal of fluoride, nitrate, and calcium by Pseudomonas sp. WZ39: Mechanisms and nucleation pathways. J. Hazard. Mater. 2021, 416, 125914. [Google Scholar] [CrossRef]
  12. Mousazadeh, M.; Alizadeh, S.M.; Frontistis, Z.; Kabdasli, I.; Niaragh, E.K.; Qodah, Z.A.; Naghdali, Z.; Mahmoud, A.E.D.; Sandoval, M.A.; Butler, E.; et al. Electrocoagulation as a promising defluoridation technology from water: A review of state of the art of removal mechanisms and performance trends. Water 2021, 13, 656. [Google Scholar] [CrossRef]
  13. Lodi, M.B.; Fanari, F.; Fanti, A.; Desogus, F.; Getaneh, W.; Mazzarella, G.; Valera, P. Preliminary study and numerical investigation of an electrostatic unit for the removal of fluoride from thermal water of ethiopian rift valley. IEEE J. Multiscale Multiphys. Comput. Tech. 2020, 5, 72–82. [Google Scholar] [CrossRef]
  14. Wioeniewski, J.; Rozanska, A. Donnan dialysis for hardness removal from water before electrodialytic desalination. Desalination 2007, 212, 251–260. [Google Scholar] [CrossRef]
  15. Nemade, P.D.; Vasudeva, R.A.; Alappadt, B.J. Removal of fluorides from water using low cost adsorbents. Water Sci. Tech.-Water Supply 2002, 2, 311–317. [Google Scholar] [CrossRef]
  16. Mohapatra, M.; Anand, S.; Mishra, B.K.; Giles, D.E.; Singh, P. Review of fluoride removal from drinking water. J. Environ. Manag. 2009, 91, 67–77. [Google Scholar] [CrossRef] [PubMed]
  17. Kaseva, M.E. Optimization of regenerated bone char for fluoride removal in drinking water: A case study in Tanzania. J. Water Health 2006, 4, 139–147. [Google Scholar] [CrossRef] [Green Version]
  18. Viswanathan, N.; Sundaram, C.S.; Meenakshi, S. Sorption behavior of fluoride on carboxylated cross–linked chitosan. J. Colloids Surf. B. 2009, 68, 48–54. [Google Scholar] [CrossRef]
  19. Paudyal, H.; Pangeni, B.; Inoue, K.; Kawakita, H.; Ohto, K.; Harada, H.; Alam, S. Adsorptive removal of fluoride from aqueous solution using orange waste loaded with multi–valent metal ions. J. Hazard. Mater. 2011, 192, 676–682. [Google Scholar] [CrossRef]
  20. Wang, X.H.; Song, R.H.; Yang, H.C.; Shi, Y.J.; Dang, G.B.; Yang, S.; Zhao, Y. Fluoride adsorption on carboxylated aerobic granules containing Ce(III). Bioresour. Technol. 2013, 127, 106–111. [Google Scholar] [CrossRef]
  21. Wang, K.; Feng, J.; Xie, S.L. Phytoremediation of phenol using Polygonum orientale, including optimized conditions. Environ. Monit. Assess. 2014, 186, 8667–8681. [Google Scholar] [CrossRef] [PubMed]
  22. Zhang, J.; Shi, Q.Q.; Zhang, C.L.; Xu, J.T.; Zhai, B.; Zhang, B. Adsorption of neutral red onto Mn–impregnated activated carbons prepared from Typha orientalis. Bioresource Technol. 2008, 99, 8974–8980. [Google Scholar] [CrossRef] [PubMed]
  23. Shi, Q.Q.; Zhang, J.; Zhang, C.L.; Bie, W.; Zhang, B.; Zhang, H.Y. Adsorption of basic violet 14 in aqueous solution using KMnO4–modified activated carbon. J. Colloid Interface Sci. 2010, 343, 188–193. [Google Scholar] [CrossRef] [PubMed]
  24. Feng, J.; Shi, S.L.; Pei, L.Y.; Lv, J.P.; Liu, Q.; Xie, S.L. Preparation of activated carbon from Polygonum orientale Linn. to remove the phenol in aqueous solutions. PLoS ONE 2016, 11, e0164744. [Google Scholar] [CrossRef] [PubMed]
  25. Jagtap, S.; Yenkie, M.K.; Labhsetwar, N.; Rayalu, S. Fluoride in drinking water and defluoridation of water. Chem. Rev. 2012, 112, 2454–2466. [Google Scholar] [CrossRef]
  26. Wang, L.; Zhang, J.; Zhao, R.; Li, Y.; Zhang, C.L.; Li, C.; Li, Y. Adsorption of 2,4–dichlorophenol on Mn–modified activated carbon prepared from Polygonum orientale Linn. Desalination 2011, 266, 175–181. [Google Scholar] [CrossRef]
  27. Feng, J.; Qiao, K.; Pei, L.Y.; Lv, J.P.; Xie, S.L. Using activated carbon prepared from Typha orientalis Presl to remove phenol from aqueous solutions. Ecol. Eng. 2015, 84, 209–217. [Google Scholar] [CrossRef]
  28. Ku, Y.; Chiou, H.M. The adsorption of fluoride ion from aqueous solution by activated alumina. Water Air Soil Poll. 2002, 133, 349–361. [Google Scholar] [CrossRef]
  29. Farrah, H.; Slavek, J.; Pickering, W.F. Fluoride interactions with hydrous aluminum oxides and alumina. Aust. J. Soil Res. 1987, 25, 55–69. [Google Scholar] [CrossRef]
  30. Wasay, S.A.; Tokunaga, S.; Park, S.W. Removal of hazardous anions from aqueous solutions by La (III)– and Y(lll)–impregnated alumina. Sep. Sci. Technol. 1996, 31, 1501–1514. [Google Scholar] [CrossRef]
  31. Maliyekkal, S.M.; Sharma, A.K.; Philip, L. Manganese–oxide–coated alumina: A promising sorbent for defluoridation of water. Water Res. 2006, 40, 3497–3506. [Google Scholar] [CrossRef] [PubMed]
  32. Puri, B.K.; Balani, S. Trace determination of fluoride using lanthanum hydroxide supported on alumina. J. Environ. Sci. Heal. A 2000, 35, 109–121. [Google Scholar] [CrossRef]
  33. Kamble, S.P.; Jagtap, S.; Labhsetwar, N.K.; Thakare, D.; Godfrey, S.; Devotta, S.; Rayalu, S.S. Defluoridation of drinking water using chitin, chitosan and lanthanum–modified chitosan. Chem. Eng. J. 2007, 129, 173–180. [Google Scholar] [CrossRef]
  34. Parmar, H.S.; Patel, J.B.; Sudhakar, P.; Koshy, V.J. Removal of fluoride from water with powdered corn cobs. J. Environ. Sci. Eng. 2006, 48, 135–138. [Google Scholar]
  35. Mohan, D.; Singh, K.P.; Singh, V.K. Wastewater treatment using low cost activated carbons derived from agricultural by products–A case study. J. Hazard. Mater. 2008, 152, 1045–1053. [Google Scholar] [CrossRef]
  36. 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]
  37. Weber, T.W.; Chakkravorti, R.K. Pore and solid diffusion models for fixed bed adsorbers. Aiche J. 1974, 20, 228. [Google Scholar] [CrossRef]
  38. Freundlich, H. Adsorption in solution. Phy. Chem. Soc. 1906, 40, 1361–1368. [Google Scholar]
  39. Kalavathy, M.H.; Karthikeyan, T.; Rajgopal, S. Kinetic and isotherm studies of Cu (II) adsorption onto H3PO4–activated rubber wood sawdust. J. Colloid Interface Sci. 2005, 292, 354–362. [Google Scholar] [CrossRef]
  40. Ho, Y.S.; McKay, G. Pseudo–second order model for sorption processes. Process Biochem. 1999, 34, 451–465. [Google Scholar] [CrossRef]
  41. Pavan, F.; Dias, S.; Lima, E.; Benvenutti, E. Removal of Congo red from aqueous solution by anilinepropylsilica xerogel. J. Dye. Pigment. 2008, 76, 64–69. [Google Scholar] [CrossRef]
  42. Kusrini, E.; Sofyan, N.; Suwartha, N.; Yesya, G.; Priadi, C.R. Chitosan-praseodymium complex for adsorption of fluoride ions from water. J. Rare Earth. 2015, 33, 1104–1113. [Google Scholar] [CrossRef]
  43. Biswas, K.; Gupta, K.; Chand Ghosh, U. Adsorption of fluoride by hydrous iron(III)–tin(IV) bimetal mixed oxide from the aqueous solutions. Chem. Eng. J. 2009, 149, 196–206. [Google Scholar] [CrossRef]
  44. Kumarb, E.; Bhatnagara, A.; Kumarc, U.; Sillanpaad, M. Defluoridation from aqueous solutions by nano-alumina: Characterization and sorption studies. J. Hazard. Mater. 2011, 186, 1042–1049. [Google Scholar] [CrossRef] [PubMed]
  45. Kemer, B.; Ozdes, D.; Gundogdu, A.; Bulut, V.N.; Duran, C.; Soylak, M. Removal of fluoride ions from aqueous solution by waste mud. J. Hazard. Mater. 2009, 168, 888–894. [Google Scholar] [CrossRef]
  46. Liang, P.; Zhang, Y.; Wang, D.F.; Xu, Y.; Luo, L. Preparation of mixed rare earths modified chitosan for fluoride adsorption. J. Rare Earth. 2013, 31, 817–822. [Google Scholar] [CrossRef]
  47. Dehghani, M.H.; Haghighat, G.A.; Yetilmezsoy, K.; McKay, G.; Heibati, B.; Tyagie, I.; Agarwal, S.; Gupta, V.K. Adsorptive removal of fluoride from aqueous solution using single- andmulti-walled carbon nanotubes. J. Mol. Liq. 2016, 216, 401–410. [Google Scholar] [CrossRef]
  48. Sun, Y.; Fang, Q.; Dong, J.; Cheng, X.; Xu, J. Removal of fluoride from drinking water by natural stilbite zeolite modified with Fe (III). Desalination 2011, 277, 121–127. [Google Scholar] [CrossRef]
  49. Chen, N.; Zhang, Z.H.; Feng, C.H.; Li, M.; Zhu, D.; Chen, R.; Sugiur, N. An excellent fluoride sorption behavior of ceramic adsorbent. J. Hazard. Mater. 2010, 183, 460–465. [Google Scholar] [CrossRef]
  50. Dehghani, M.H.; Farhang, M.; Alimohammadi, M.; Afsharnia, M.; Mckay, G. Adsorptive removal of fluoride from water by activated carbon derived from CaCl2-modified Crocus sativus leaves: Equilibrium adsorption isotherms, optimization, and influence of anions. Chem. Eng. Comm. 2018, 205, 955–965. [Google Scholar] [CrossRef]
  51. Gao, Y.N.; Li, M.F.; Ru, Y.F.; Fu, J.X. Fluoride removal from water by using micron zirconia/zeolite molecular sieve: Characterization and mechanism. Groundw. Sustain. Dev. 2021, 13, 100567. [Google Scholar] [CrossRef]
Figure 1. Pseudo-first-order results for fluoride adsorption in (a) PO, (b) PO–Al, and (c) PO–La systems.
Figure 1. Pseudo-first-order results for fluoride adsorption in (a) PO, (b) PO–Al, and (c) PO–La systems.
Water 14 00592 g001
Figure 2. Pseudo-second-order results for fluoride adsorption in (a) PO, (b) PO–Al, and (c) PO–La systems.
Figure 2. Pseudo-second-order results for fluoride adsorption in (a) PO, (b) PO–Al, and (c) PO–La systems.
Water 14 00592 g002
Figure 3. Langmuir isotherms for fluoride adsorption onto (a) PO, (b) PO–Al, and (c) PO–La at different temperatures.
Figure 3. Langmuir isotherms for fluoride adsorption onto (a) PO, (b) PO–Al, and (c) PO–La at different temperatures.
Water 14 00592 g003
Figure 4. Freundlich isotherms for fluoride adsorption onto (a) PO, (b) PO–Al, and (c) PO–La at different temperatures.
Figure 4. Freundlich isotherms for fluoride adsorption onto (a) PO, (b) PO–Al, and (c) PO–La at different temperatures.
Water 14 00592 g004
Figure 5. Resorption of fluoride on PO, PO–Al, PO–La desorbed by NaOH, HCl, deionized water, and ethanol.
Figure 5. Resorption of fluoride on PO, PO–Al, PO–La desorbed by NaOH, HCl, deionized water, and ethanol.
Water 14 00592 g005
Table 1. Pseudo-first-order and pseudo-second-order kinetic parameters.
Table 1. Pseudo-first-order and pseudo-second-order kinetic parameters.
Pseudo-First-OrderPseudo-Second-Order
Ce
(mg/L)
qe (exp)
(mg/g)
k1 × 10−2
(min−1)
qe (cal)
(mg/g)
R2k2
(g/mg·min)
qe
(cal) (mg/g)
h
(mg/g·min)
R2
PO
100.29350.01040.09930.91110.39700.29630.03490.9962
200.64690.01310.13260.98080.36150.65250.15390.9993
300.40750.00990.34350.85850.02760.52080.00750.9609
PO-Al
100.47780.01360.07790.95640.59470.48290.13870.9994
200.94320.09560.04150.90311.27700.94451.13910.9996
300.07580.03290.30160.98050.20701.42550.42070.9994
PO-La
100.39800.01910.28120.76200.18740.39800.02970.7445
200.79670.01360.27070.96720.21960.78440.13510.9056
300.56470.00600.40690.88100.12660.42090.02240.9933
Table 2. Langmuir and Freundlich isotherm parameters for fluoride adsorption in PO, PO–Al and PO–La systems.
Table 2. Langmuir and Freundlich isotherm parameters for fluoride adsorption in PO, PO–Al and PO–La systems.
T (K)LangmuirFreundlich
Qm (mg/g)b (L/mg)RLR2KF (mg/g (L/mg)1/n) 1/nR2
PO
2830.09160.22110 < RL < 10.86410.97140.66310.6577
2930.40420.73310 < RL < 10.90490.39890.10630.0341
3030.09230.17580 < RL < 10.80941.02140.5390.4606
PO–Al
2830.70605.42510 < RL < 10.86870.90411.41150.7420
2930.84033.09120 < RL < 10.94620.77021.38380.9477
3033.31900.41140 < RL < 10.71080.95060.73060.9028
PO–La
2830.57352.02910 < RL < 10.96440.33980.2060.5367
2930.66451.17170 < RL < 10.96860.53980.30860.4247
3031.25140.23780 < RL < 10.98900.30160.47580.9328
Table 3. Comparison of the fluoride adsorption capacity in this study and reported in the literature.
Table 3. Comparison of the fluoride adsorption capacity in this study and reported in the literature.
AdsorbentsAdsorption Capacity (mg/g)References
Chi-Pr8.20[42]
Hydrous iron(III)–tin(IV) bimetal10[43]
Nano-alumina14[44]
o-WM4.2[45]
CR3.72[46]
CL3.16[46]
MOCA2.85[31]
MWCNTs2.83[47]
a-WM2.8[45]
SWCNTs2.40[47]
Fe(III)-STI2.31[48]
Ceramic2.16[49]
AC-CMCSL2.01[50]
ZrO2-Ze0.34[51]
PO-Al3.32This study
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Shi, S.; Zhao, K.; Zhao, Q.; Luo, A.; Xie, S.; Feng, J. Fluoride Adsorption Comparison from Aqueous Solutions Using Al- and La-Modified Adsorbent Prepared from Polygonum orientale Linn. Water 2022, 14, 592. https://doi.org/10.3390/w14040592

AMA Style

Shi S, Zhao K, Zhao Q, Luo A, Xie S, Feng J. Fluoride Adsorption Comparison from Aqueous Solutions Using Al- and La-Modified Adsorbent Prepared from Polygonum orientale Linn. Water. 2022; 14(4):592. https://doi.org/10.3390/w14040592

Chicago/Turabian Style

Shi, Shengli, Kangxu Zhao, Qiyue Zhao, Aiguo Luo, Shulian Xie, and Jia Feng. 2022. "Fluoride Adsorption Comparison from Aqueous Solutions Using Al- and La-Modified Adsorbent Prepared from Polygonum orientale Linn." Water 14, no. 4: 592. https://doi.org/10.3390/w14040592

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