Next Article in Journal
Validation of Antibacterial Systems for Sustainable Ceramic Tiles
Previous Article in Journal
The Effect of Gaseous Ozone and Moringa Leaf–Carboxymethyl Cellulose Edible Coating on Antioxidant Activity and Biochemical Properties of ‘Keitt’ Mango Fruit
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Arsenic Removal from Contaminated Water Using Natural Adsorbents: A Review

1
School of Human Settlements and Civil Engineering, Xi’an Jiaotong University, Xi’an 710049, China
2
School of Environmental and Municipal Engineering, Xi’an University of Architecture and Technology, Xi’an 710055, China
*
Author to whom correspondence should be addressed.
Coatings 2021, 11(11), 1407; https://doi.org/10.3390/coatings11111407
Submission received: 12 October 2021 / Revised: 2 November 2021 / Accepted: 3 November 2021 / Published: 19 November 2021

Abstract

:
More than 170 million individuals have been influenced by arsenic (As) because of the ingestion of As-polluted groundwater. The presence of As in water bodies, particularly groundwater, has been found to become a widespread issue in the past few decades. Because arsenic causes extreme wellbeing impacts, even at a low concentration in drinking water, the innovations of As removal from contaminated water are of significant importance. Traditional strategies, for example, reverse osmosis, ion exchange, and electro-dialysis are generally utilized for the remediation of As-polluted water; however, the high cost and/or sludge production restricts their application in less-developed areas. The utilization of adsorbents acquired from natural materials has been explored as an alternative for the costly techniques for As removal. This paper aims to review the past and current developments in using naturals adsorbents or modified natural materials for arsenic removal and show the different parameters, which may influence the As removal effectiveness of the natural adsorbent, such as contact time, adsorbent dosage, flow rate, pH, reusability, temperature, and influence of others ions.

1. Introduction

The name arsenic (As) is thought to originate from the Syriac, Arabic, or Persian languages, and was later adopted into the Greek word arsenikon. The yellow-colored orpiment (As2S3) is infamous as the lord of all poisons. Arsenic (As) is an exceptionally harmful and cancer-causing metalloid that accounts for approximately 0.00005% of the Earth’s outside layer. As ranks as the twentieth among the elements in abundance in the Earth’s crust, fourteenth in the seawater, and twelfth in human body [1]. Arsenic occurs in rock soils, organisms, natural water, air, and in the other natural environments [2]. The element of As can form over 200 minerals in soil, and mainly exists as As(0), As(III), As(V), and arsenide. Besides, a high As concentration is found in numerous oxide minerals and hydrous metal oxides, either as a part of the mineral structure or as adsorbed species. The amount of As in sedimentary rocks ordinarily ranges from 5 to 10 mg/kg, which is usually over the normal content in a terrestrial environment, approximately 1.5 to 3 mg/kg [3]. A significant amount of As in natural water is also from anthropogenic sources, such as cosmetics, mining operations, coal combustion, smelting, pesticides, dyes, wood treatment, As trioxide stockpiles and paints. Arsenic is also highly soluble in water [4,5,6].
The two most prevailing types of As in aqueous environment are As(III) and As(V). Depending on the environment conditions, such as pH, the redox condition, the microbial activity, and presence of complexing ions, the pentavalent As mainly consists of arsenate ions (H2AsO41−, HAsO42−, AsO43−) and arsenic acid (H3AsO4). Similarly, the As(III) species may present as arsenite ions (H2AsO31−, HAsO32−, and AsO33−) and arsenious acid (H3AsO3). The pentavalent species are the predominant species under aerobic conditions, such as surface water [7], whereas the trivalent species are the prevailing species under anoxic conditions, such as subsurface groundwater [8,9].

2. Harmful Effects of Arsenic

As(III) and As(V) are both very toxic, with increased risk for cancer of the skin, lungs, urinary bladder, liver, and kidney [10], but As(III) is multiple times more harmful and versatile than As(V). Wellbeing impacts of arsenic rely upon different kind of parameters. These include the type and amount of arsenic that has been swallowed, how long you have been exposed to arsenic, and how the body reacts to arsenic. Unborn infants, children, and people with long sicknesses and the elderly are the weakest to arsenic exposure. The impacts of arsenic, when ingested even in limited quantities, show up gradually; it might take quite a long while for the harming effects to become obvious.
Arsenic pollution in natural water is an overall issue, and has become a significant issue and challenge for world specialists, researchers, and even politicians. For instance, persistent arsenic poisoning due to drinking arsenic-polluted water has been one of the terrible wellbeing impacts influencing eight areas of West Bengal since the mid 1980s. Itemized clinical assessment and investigation of 248 such patients showed the clinical signs of such harmfulness. Well beyond hyperpigmentation and keratosis, powerlessness, paleness, painful eyes, leg swelling, liver fibrosis, persistent lung ailment, gangrene of toes, neuropathy, and skin malignant growth were a portion of different symptoms [11]. The World Health Organization (WHO) and the U.S. Environmental Protection Agency (US EPA) have limited the As concentration level at 10 μg/L in 2001.

3. Arsenic Removal Technology

Considering the danger of As in water, especially in drinking water, many approaches have been studied for As removal. Traditionally, there are a few techniques including membrane filtration, coagulation and flocculation, electrochemical techniques, bioremediation, ion or particle exchange, and adsorption utilized in arsenic removal [12,13,14,15].

3.1. Membrane Filtration

Membrane filtration is most commonly used to eliminate arsenic from water. Membrane films are regularly manufactured materials with billions of pores or minute openings that operate as a specific obstruction; the structure of membrane permits a few constituents to go through, while others are avoided or dismissed. The different kinds of membrane filtration can be divided into four techniques: ultrafiltration membranes are usually with 0.0003 to 0.1 microns for colloids, viruses, and certain proteins removal [16]; nanofiltration with 0.001 to 0.003 microns for physical removal based on both charge and molecular size; microfiltration with 0.1 micron for bacteria and suspended solids removal [17]; and hyperfiltration or reverse osmosis with 0.0005 microns for desalinization [18].

3.2. Coagulation

Many studies have showed that coagulation can be used for removing arsenic from contaminated water by the addition of a coagulant in the arsenic solution. This process can also remove many other suspended and dissolved constituents from water, such as fluoride, manganese, iron, phosphate, and turbidity. This arsenic removal technique is profoundly reliant upon pH, the valence of the arsenic species., dosage of coagulant, and the initial arsenic concentration [19].

3.3. Electrochemical Techniques

The utilization of electricity for waste water treatment started early, since 1887 in the United Kingdom [20]. There are a lot of different electrochemical techniques, including electrodeposition [21], electrocoagulation [22,23], electrodialysis [24], electroflotation [25], and electrooxidation [26] used for heavy metal removal. Electrocoagulation is a process using chemical and physical reactions in which sacrificial electrodes are used for the generation of coagulants. This process, needing a high amount of electricity, is the one used the most for arsenic removal from water.

3.4. Bioremediation

Some natural organisms, such as bacteria, algae, and plants especially the aquatic plants, have a capacity to amass trace elements (heavy metals), which allows them to be used for the treatment of polluted water from industrial or agrochemical discharges, especially contaminated with heavy metals [27,28].

3.5. Ion or Particle Exchange

Ion exchange is a physical/chemical process in which particles held electrostatically on the outside of a solid phase are exchanged for particles of same charge in a solution (i.e., drinking water). The solid is ordinarily a synthetic or natural anion exchange resin, which is utilized to specially remove specific contaminants of concern. Ion exchange is generally utilized in drinking water treatment for softening (i.e., removal of magnesium, calcium, and other cations in exchange of sodium), as well as removing arsenate, nitrate, selenate, and chromate from municipal water [29].

3.6. Adsorption

Adsorption is a process that utilizes solids for eliminating substances from either liquid or gaseous solutions, by attraction on the solid’s surface. Solids, such as synthetic materials (fibers, resins, activated carbon, metal hydrides, membranes) [30,31] or natural materials [32] are widely utilized in industrial application for wastewater treatment and purification. However, synthetic materials are expensive to be produced, as the production of them requires a lot of energy and is not environment friendly compared to natural materials.
As mentioned above, most processes have been proved to be effective in removing As from water. For example, As in water can be effectively removed after being blended with other metal oxides (iron or aluminum oxides) for a coagulation to change into the solid phase [33]. However, the high energy required (electrocoagulation), the high cost of their process (membrane filtration), and the non-environmentally friendly residue (coagulation and synthetic adsorbent adsorption) after the treatment processes, limit their application and lead researchers to conduct new research with naturals materials, which are ecological and cheaper to find [34,35,36]. In this paper, we reviewed the removal performances of As from contaminated drinking water/groundwater by adsorption using natural adsorbents and/or modified natural materials.

4. Biosorbents

Biosorbents (natural adsorbents) are made from natural materials, including natural fibers, volcanic rocks, soils, plant biomass, agricultural and industrial wastes, animal shells, microalgae, and fungal biomass (Figure 1). A common feature of these materials is that their large specific surface area, which is capable of physically retaining arsenic ions or molecules. In recent years, more and more natural adsorbents are used because of their higher arsenic removal capacity, reusability, and lower cost and environment impacts as compared with the other methods involving synthetics membranes and materials using significant chemical dosage [37,38]. Among all the natural materials, agricultural wastes, comprised of proteins, extractives, hemicellulose, lignin, lipids, starch, and simple sugars, contain more functional groups, which facilitate heavy metal retention [39,40,41,42]. In Table 1, some different natural adsorbents that have been used for As removal from water are listed.

5. Treatment of the Natural Adsorbents

Some natural adsorbents without any kind of pre-treatment are good enough for removing As from water because of their surface structure composition, such as the acidic functional groups (–SH, –COOH) present in the rice husk cellulose component or the Fe and Al oxide present on montmorillonite, kaolinite, and illite clay minerals as impurities (Figure 2). These functional groups may facilitate the interaction with As anions [68,69]. On the other hand, as we can see in the Table 1, some adsorbents need to be pre-treated to improve their As adsorption capacity. The goal of pre-treatment by washing and cleaning, using distilled water, alkali (NaOH) or acidic (HCl) solution is to remove impurity particles or add some functional group on the surface of the natural material. Therefore, pre-treatment will make it possible to highlight the functional groups (CH3–OH, –COOH, CH3–NH2, Fe(OH)3, Al(OH)3), and therefore will facilitated the interactivity between natural adsorbents and As ions.
Distilled water is used most of the time to remove surface adhered impurity particles, water-soluble materials [57]. The distilled water washed materials will be dried under specific conditions as necessary, for example, coconut coir pith (CP) dried at 80 °C, groundnut shell dried under sun for 2 days [58], and rice husk dried at 60 °C [70]. Alkali pre-treatment with 20% NaOH solution is used to degrease cotton and add hydroxyl group onto the surface [59], and remove the lignin of sawdust [60]. Acidic pre-treatment selectively removes the organic base by converting it into a water-soluble salt as HCL (10%), which is used to clean rice husk to remove all the impurities, then later crushed and dried at 500 °C in muffle furnace for 8 h [61], and Dialium guineense seed shells have been crushed then soaked in (40%) phosphoric and nitric acid, then heated to a carbon activation temperature of 400 °C for 30 min.
The objective of the treatment is to change the surface structure, to improve the As removal capacity of natural adsorbents by adding essential functional groups (Fe3+, Al3+, TiO2) [71,72,73,74]. Aluminum ions are often used for the modification of the natural adsorbent’s surface, since Al3+ oxidative capacity towards As(III) is well known [75,76]. The reactions below might be the interaction of As(V) and As(III) with the natural asorbent (M) after its surface modification with aluminum:
As(V)
M=Al(OH) + H3AsO4 → M=Al–AsO4H2 + H2O
M=Al(OH) + H2AsO4 → M=Al–HAsO4 + H2O
M=Al(OH) + HAsO42− → M=Al–AsO42− + H2O
As(III)
M=Al(OH) + H3AsO3 → M=Al–AsO3H2 + H2O
M=Al(OH) + H2AsO3 → M=Al–HAsO3 + H2O
M=Al(OH) + HAsO32− → M=Al–AsO32− + H2O
For example, hydroxyl-alumina has been used as a surface activating agent for the modification of paddy husk ash particles, using AlCl3·6H2O and aluminum fine powder [62]. The surface of crushed pumice and zeolite stone particles can be coated with aluminum ions by immersing pumice and zeolite stone particles in 0.5 M Al2(SO4)·16H2O solution, and drying at room temperature for 72 h [63]. Similarly, the removal of As(V) and As(III) can be enhanced with the introduction of ferric ions on the surface of natural materials (M) because of the formation of ferric arsenate FeAsO4·2H2O at a low pH [67,77,78]. The reactions below might be the interaction of As(V) and As(III) with the natural adsorbent (M) after its surface modification with iron:
As(V)
M=Fe(OH) + H3AsO4 → M=Fe–AsO4H2 + H2O
M=Fe(OH) + H2AsO4→M=Fe–HAsO4 + H2O
M=Fe(OH) + HAsO42− → M=Fe–AsO42− + H2O
As(III)
M=Fe(OH) + H3AsO3 → M=Fe–AsO3H2 + H2O
M=Fe(OH) + H2AsO3 → M=Fe–HAsO3 + H2O
M=Fe(OH) + HAsO32− → M=Fe–AsO32− + H2O
Maji et al. [79] and Joshi et al. [80] found that the surface structure of crushed natural rock particles and river sand could be successfully modified using Fe(III) nitrate solution, then dried in 110–150 °C hot air; this finding was similar to Jeon C.S. et al. [66] using zeolite (clinoptilolite) as a natural material. In some cases, natural iron oxide minerals, such as magnetite (FeOFe2O3), hematite (Fe2O3), goethite (FeO(OH)), and laterite, were used directly for As adsorption from water without any pre-treatment or treatment [65].

6. As Removal Performance

Many researches have been conducted to evaluate the As adsorption performance of many different types of natural adsorbents, by studying the effects of different factors [81]. There are many parameters that can influence the removal capacity of the natural adsorbents, such as pH, initial dosage of the adsorbents, initial concentration of the As solution, contact time, temperature, and the effect of other ions in the water solution. The natural adsorbent for As removal has mainly been tested by using two different methods, batch test and column test [82,83,84,85,86].

6.1. Overall Adsorption Performance

To achieve an efficient As removal from water, both the initial As concentration and adsorbent dosage are optimized before conduction water treatment [87,88]. As can be seen in the Table 2, it is obvious that an increase in the initial As concentration leads to a decrease in the As removal efficiency [89,90]. Usually, As removal efficiency increases with the increase in the adsorbent’s dosage, due to the higher surface area or exchangeable sites provided. However, if the adsorbent’s dosage is too high, the removal efficiency of As will finally stay constant, which is not economic [91,92,93,94]. Take iron-modified peat for example, it was determined that only 60% of the As(V) was removed with an initial concentration of 800 mgžL−1, and the ratio reached more than 90% when its initial concentration decreased to 300 mgžL−1 [95,96]. Compared with iron-modified peat, PAC-500 (peels activated carbon) and PPAC-500 (pulps activated carbon) seem to have a lower As adsorption capacity. It was determined that when the initial concentration of As was around 2.5 mg/L, less than 80% of As was removed, and this ratio would be above 95% when the initial concentration of As decreased to around 0.5 mg/L [97]. Again, considering only 40–50% of As was adsorbed when its initial concentration was in the range of 0.005–0.02 mg/L, modified clinoptilolite zeolite (MCZ) seems have the lowest As adsorption capacity [98].

6.2. Effects of Contact Time on As Adsorption

The evaluation of operational parameters, such as the contact time (batch adsorption) and flow rate (column adsorption) are also important to determine the time at which the adsorption reaches equilibrium. The quantity of As adsorbed per unit mass of adsorbent, and the percentage of As adsorbed will be determined respectively using Equations (13) and (14) given below:
qt = ((C0Ct)/M) × V
A(%) = ((C0Ct)/C0) × 100
where, qt is the quantity of As adsorbed at time t; A(%) is the percentage of As adsorbed at time t; C0 is the initial As concentration; Ct is the concentration of As in the aqueous phase at time t; V is the volume of the aqueous solution; and M is the mass of natural As that will be used in the experiment [102].
The increase in the flow rate causes a decrease in the contact time between the adsorbent and As solution, which in turn lowers its removal efficiency. If the contact time is not long enough, the surface of the adsorbent will not be able to be charged enough, which will be a waste of the adsorbent. On the other side, if the contact time is longer than the equilibrium time, the adsorbent might start releasing back some As particles in the aqueous solution, which will not benefit the removal efficiency [103,104]. Frequently, the quantity of As adsorbed increases rapidly because there are still many adsorption sites on the natural adsorbent surface at first, then it will gradually slow down until it stays constant after the equilibrium adsorption time [105,106,107,108,109].
The augmentation of flow rate on As adsorption by thioglycolated sugarcane carbon (TSCC) and iron-modified cement slurry have been studied. The collected data showed that the As removal performance decreased with the increase in the flow rate from 3.0 to 7.0 mL min−1 and 5 to 14 mL/min [103,110]. However, the results using (MNCF) modified nonwoven cotton fabric, modified rice husk particles for As removal from water, showed that within 1 h a high As adsorption rate (almost 80%) was noticed [55,111]. The marine sand might have fast As adsorption in comparison to algae, as zirconium oxide-coated marine sand (ZrOCMS) removed more than 90% within 75 min, which is quick [112] compared to bio-adsorbents, such as algae (Lessonia nigrescens), in which the maximum adsorption capacity was reached after 300 min [113].

6.3. pH

The adsorption of As ions species are strongly governed by the pH of the aqueous solution. The pH of the solution is an important parameter, which controls the As species present in water and the natural adsorbent surface chemical composition [114]. The distributing equation of As(III) is as follows:
H3AsO3 = H+ + H2AsO3, pKa1 = 9.2
H2AsO3 = H+ + HAsO32−, pKa2 = 12.1
HAsO32− = H+ + AsO33−, pKa3 = 13.4
The distributing equation of As(V) is as follows:
H3AsO4 = H+ + H2AsO4, pKa1 = 2.1
H2AsO4 = H+ + HAsO42−, pKa2 = 7
HAsO42− = H+ + AsO43−, pKa3 = 11.2
Mostly, if the As removal is through chemical adsorption, the As(III) adsorption is better at a high pH [115] because at pH 1–7 the predominant As(III) species H3AsO3 (pKa = 9.2) is uncharged, which can negatively impact the As removal performance. As shown in Figure 3, the adsorption performance of As(III) increased notably with the increase in the solution pH, and the best solution pH for As(III) adsorption was around pH = 8.5 when the copper impregnated coconut husk carbon (CICHC) was used as the adsorbent [116]. A study using fly ash from the power plant after burning biomass and coal for As(III) removal showed that the best removal data was around pH 12 [117]. However, for As(V), its removal is mostly better at a low pH because the adsorption performance decreases with the diminution of H2AsO4 (pKa = 7) percentage in the aqueous solution [81,118,119,120,121,122]. As it can be seen in the Figure 4, Wang et al. [123], using Ni/Fe modified loblolly pine (Pinus taeda) wood biochar (NFMB) for As(V) removal, showed that the adsorption capacity decreased with the solution pH rising from 3 to 9. Another study using iron-modified loblolly pine (Pinus taeda) wood biochar (nZVI/BC) for As(V) removal showed that the removal efficiency decreased briskly at pHs 3–3.7, and kept dropping as the solution pH increased [124].

6.4. Reusability

Desorption and regeneration are important parameters because a good natural adsorbent is an adsorbent that can remove enough As from water and can be reused without losing its adsorption performance too much, which indicates that the adsorbent can be recycled and reused. The loss of the adsorption capacity, after much desorption and regeneration, is due to the loss of adsorption sites on the adsorbent surface [90,123,125,126]. Generally, it can be noticed in the Table 3, most of the natural adsorbents have good reuse performance. It was determined that after the iron hydroxide/manganese dioxide doped straw activated carbon (Fe-Mn-Sac) was used for three adsorption–desorption cycles, the adsorption capacity of Fe-Mn-Sac had a negligible variation (from 85% to 78%) [127]. For iron-modified water hyacinth biochar, the arsenate removal percentage decreased from 100 to 65% after four cycles of regeneration, which was not a bad performance regarding the high arsenate initial concentration in the water (5 mg/L) [128].

6.5. Thermodynamic

The variation in temperature can have positive or negative impacts on the adsorbent’s adsorption capacity [134,135]. Usually, for the adsorption thermodynamic study, Gibbs free energy (ΔG°), entropy (ΔS°), and enthalpy (ΔH°) of the adsorption are calculated using van’t Hoff thermo-dynamic equations:
ΔG° = −RT ln(KD)
ln(KD) = ΔS°/R − ΔH°/(RT)
where R is the universal gas constant (8.314 J/mol K), T is temperature (K), and KD (qe/Ce) is the distribution coefficient. According to Equation (22), the entropy (ΔS°) and enthalpy (ΔH°) of the adsorption parameters can be determined from the slope and intercept of the plot of ln(KD) vs. 1/T yields, respectively, and then used to calculate the Gibbs free energy (ΔG°) [129].
The term physical sorption means a van der Waals type force formed between the interfaces and chemisorption, which denotes a chemical bond formed between the As molecule and the natural adsorbent surface. Generally, the physical sorption enthalpy (ΔH°) is in the range of −20 to −40 kJ mol−1 and chemisorption enthalpy (ΔH°) in the range of −400 to −80 kJ mol−1 [136,137]. Usually, the adsorption of As on natural adsorbents are spontaneous chemisorption processes [133]. Take chemically modified watermelon rind for example, the negative values of ΔG° suggested that both As(III) and As(V) adsorption was spontaneous; the positive values of ΔH° for both As(III) (67 kJ mol−1) and As(V) (86.05 kJ mol−1) sorption indicated that sorption was endothermic; the positive ΔS° values for both As(III) (0.24 kJ mol−1) and As(V) (0.58 kJ mol−1) suggested a disorder at the solid/solution interface along with structural changes [132].

6.6. Influence of Other Ions

Many common ions which can be present in water, especially groundwater, can negatively impact the removal of As due to the capacity of these ions to compete with As for adsorption [138]. In general, as shown in the Table 4, the adsorption capacity of As onto a natural adsorbent in the presence of positive anions (Mg2+, Ca2+, Mn2+) is higher than in the presence of negative anions (NO3, SO42−, PO43−) because of the electrostatic competition between the negative other anions and the As anions for the adsorption onto the natural adsorbent surface, and the repulsion between the positively charged natural adsorbent surface and the positive anions [89,139,140]. Minerals might have better selectivity compared to biomass, for example, leonardite was used for As removal and the results showed that the enhanced SO42− concentration in the solution led to the reduction in As removal capacity from 100 to 95% [141], while Baig et al. [135] found that the removal efficiency of As(III) on the iron modified Kans grass (Saccharum spontaneum) biochar dropped notably below 20%, when the concentration of PO43− increased to 1.0 mmol/L.

6.7. Adsorption Isotherm and Kinetic Model

In the order to design an As removal system on a real scale, studies of adsorption isotherm and kinetics equation models are important to describe the nature of adsorption onto the natural adsorbents, and determine the adsorption capacity of As [146,147,148,149]. Many isotherm equation models exist, but the most use ones in the studies of As adsorption are the Langmuir Equation (23) and Freundlich Equation (24) linear equation.
Ce/qe = Ce/qm + 1/qmKa
where Ce is the equilibrium concentration, qe is the equilibrium adsorption capacity of As onto adsorbent, qm is the maximum adsorption capacity, and Ka is the Langmuir sorption equilibrium constant [150].
logqe = 1/n × logCe + logKf
where Ce and qe have the same meaning as explained above, Kf and n are the Freundlich isotherm constants, representing the adsorption capacity and intensity, respectively [151].
The Langmuir model is an equation which is used to describe monolayer adsorption onto a homogeneous surface. The Freundlich model assumes chemisorption on heterogeneous surface [152,153]. As we can see in Table 1, the Langmuir model is the isotherm sorption model that most of the time explains the adsorption mechanism when a natural adsorbent is used.
There are two kinetics models that are mostly used, which are the pseudo-first-order model (Equation (25) [154] and pseudo-second-order model (Equation (26) [155].
log(QeQt) = logQek1/2.303t
where Qe is the amount of As adsorbed on the surface of the adsorbent at equilibrium (μg/g), Qt is amount of As on the surface of the adsorbent at time any t (μg/g), and k1 is the equilibrium rate constant of pseudo-first-order sorption (L/min). The rate constants are calculated by plotting log(QeQt) vs. t.
t/Qt = 1/k2Qe2 + t/Qe
where t, Qt, and Qe have the same meaning as explained above. k2 is the rate constant of pseudo-second-order adsorption (g·μg−1·min−1). The rate constants are calculated by plotting t/Qt vs. t.
The pseudo-second-order kinetic model assumes that chemisorption is the main adsorption mechanism [156] and the pseudo-first-order kinetic model indicates that physical sorption is the adsorption mechanism [140]. Mostly, as we can see in Table 4, when a biomaterial, such as the iron-modified chitosan hollow fibers membrane, ZnCl2-activated pig manure residue biochar, pine bark, pine wood, oak bark, and oak wood biochar is used for As(III) or As(V) removal, the principal adsorption mechanism is a chemisorption [157,158,159].

7. Conclusions

Some biosorbents, such as fruit peels (orange, citrus, mango) and minerals (geothite) have a very nice As removal performance, but some others such as biomass or fibers (cotton, sugarcane, coconut) first need a pre-treatment to enhance their performance. As removal by natural adsorbent was found to be mainly endothermic and a spontaneous monolayer chemisorption, and depended on many different parameters; high As initial concentration may reduce the biosorbent As removal effectiveness, but more contact time can help to uptake more As from the water. Enough dosage of the natural sorbent and suitable water flow are very useful for better As removal performance, the adsorption of As(V) is suitable in acidic solution and As(III) is appropriate in a basic solution. The economic factor is the main reason why it will be better to use as natural adsorbent of As from water, a natural material which has no use in our society and usually becomes waste materials, such as sugar cane fibers, corn fibers, corn cane fibers, groundnut shell. The low cost, good adsorption, and desorption of the biosorbent prove that natural adsorbents are a good As adsorbent, which can help to set up an effective technology for any kind of water polluted by As. In the future, all these advantages of the natural sorbents for As removal from water should be presented to researchers, to increase the research using biosorbents for further As adsorption technology development. Comparisons of the natural sorbent As adsorption capacity with or without modification should be made to reduce the use of chemical products in the process. Moreover, more studies should be conducted using real polluted natural water to check the performance of natural sorbent in real conditions, especially using the column process.

Author Contributions

Conceptualization, K.F.H.Y.; Formal analysis, K.F.H.Y.; Investigation, K.F.H.Y.; Writing—original draft, K.F.H.Y.; Methodology, K.F.H.Y.; Writing—review & editing, W.W.; Resources, W.W.; Supervision, W.W.; validation, C.L., H.Z., J.C., W.W. and Y.D. All authors have read and agreed to the published version of the manuscript.

Funding

Funded by Technology Innovation Center for Land Engineering and Human Settlements; Shaanxi Land Engineering Construction Group Co., Ltd.; Xi’an Jiaotong University (No. 201912131) and Science technology Project of Yulin (No. YF-2020-007).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Jain, C.; Ali, P.I. Arsenic: Occurrence, Toxicity and Speciation Techniques. Water Res. 2000, 34, 4304–4312. [Google Scholar] [CrossRef]
  2. Smedley, P.L.; Kinniburgh, D.G. A review of the source, behaviour and distribution of arsenic in natural waters. Appl. Geochem. 2002, 17, 517–568. [Google Scholar] [CrossRef] [Green Version]
  3. Mandal, B.K.; Suzuki, K.T. Arsenic round the world: A review. Talanta 2002, 58, 201–235. [Google Scholar] [CrossRef]
  4. Mohan, D.; Pittman, C.U., Jr. Arsenic removal from water/wastewater using adsorbents—A critical review. J. Hazard. Mater. 2007, 142, 1–53. [Google Scholar] [CrossRef]
  5. Naidu, R.; Smith, E.; Owens, G.; Bhattacharya, P.; Nadebaum, P.; Collingwood, V. Managing Arsenic in the Environment from Soil to Human Health; CSIRO Publishing: Clayton, Australia, 2006. [Google Scholar]
  6. Shakoor, M.B.; Niazi, N.K.; Bibi, I.; Rahman, M.M.; Naidu, R.; Dong, Z.; Shahid, M.; Arshad, M. Unraveling Health Risk and Speciation of Arsenic from Groundwater in Rural Areas of Punjab, Pakistan. Int. J. Environ. Res. Public Health 2015, 12, 12371–12390. [Google Scholar] [CrossRef] [Green Version]
  7. Cutter, G. Kinetic controls on metalloid speciation in seawater. Mar. Chem. 1992, 40, 65–80. [Google Scholar] [CrossRef]
  8. Van Genuchten, C.M.; Ahmad, A. Groundwater As Removal by As(III), Fe(II), and Mn(II) Co-Oxidation: Contrasting As Removal Pathways with O2, NaOCl, and KMnO4. Environ. Sci. Technol. 2020, 54, 15454–15464. [Google Scholar] [CrossRef]
  9. Gecol, H.; Ergican, E.; Fuchs, A. Molecular level separation of arsenic (V) from water using cationic surfactant micelles and ultrafiltration membrane. J. Membr. Sci. 2004, 241, 105–119. [Google Scholar] [CrossRef]
  10. National Research Council USA. Arsenic in Drinking Water: 2001 Update; NRC: Washington, DC, USA, 2001.
  11. Mazumder, D.N.G. Chronic Arsenic Toxicity: Clinical Features, Epidemiology, and Treatment: Experience in West Bengal. J. Environ. Sci. Health Part A 2003, 38, 141–163. [Google Scholar] [CrossRef]
  12. Babel, S.; Kurniawan, T. Low-Cost Adsorbents for Heavy Metals Uptake from Contaminated Water: A Review. J. Hazard. Mater. 2003, 97, 219–243. [Google Scholar] [CrossRef]
  13. Raul, P.; Devi, R.R.; Umlong, I.M.; Banerjee, S.; Singh, L.; Purkait, M.K. Removal of Fluoride from Water Using Iron Oxide-Hydroxide Nanoparticles. J. Nanosci. Nanotechnol. 2012, 12, 3922–3930. [Google Scholar] [CrossRef]
  14. Daus, B.; Wennrich, R.; Weiss, H. Sorption materials for arsenic removal from water: A comparative study. Water Res. 2004, 38, 2948–2954. [Google Scholar] [CrossRef]
  15. Yin, H.; Liu, F.; Feng, X.; Liu, M.; Tan, W.; Qiu, G. Co2+-exchange mechanism of birnessite and its application for the removal of Pb2+ and As(III). J. Hazard. Mater. 2011, 196, 318–326. [Google Scholar] [CrossRef]
  16. Wang, J.; Zhou, X.F.; Ma, J.X. Preparation and Characteristics of a Paper-Based Ultrafiltration Membrane. Bioresources 2012, 7, 545–553. [Google Scholar]
  17. Nguyen, V.T.; Vigneswaran, S.; Ngo, H.H.; Shon, H.K.; Kandasamy, J. Arsenic removal by a membrane hybrid filtration system. Desalination 2009, 236, 363–369. [Google Scholar] [CrossRef] [Green Version]
  18. Yang, Z.; Zhou, Y.; Feng, Z.; Rui, X.; Zhang, T.; Zhang, Z. A Review on Reverse Osmosis and Nanofiltration Membranes for Water Purification. Polymers 2019, 11, 1252. [Google Scholar] [CrossRef] [Green Version]
  19. Pallier, V.; Feuillade-Cathalifaud, G.; Serpaud, B.; Bollinger, J.-C. Effect of organic matter on arsenic removal during coagulation/flocculation treatment. J. Colloid Interface Sci. 2010, 342, 26–32. [Google Scholar] [CrossRef]
  20. Meharg, A.A.; Whitaker, J. Arsenic uptake and metabolism in arsenic resistant and nonresistant plant species. New Phytol. 2002, 154, 29–43. [Google Scholar] [CrossRef] [Green Version]
  21. Zhitomirsky, I. Cathodic electrodeposition of ceramic and organoceramic materials. Fundamental aspects. Adv. Colloid Interface Sci. 2002, 97, 279–317. [Google Scholar] [CrossRef]
  22. Mollah, M.A.; Schennach, R.; Parga, J.R.; Cocke, D.L. Electrocoagulation (EC)—Science and applications. J. Hazard. Mater. 2001, 84, 29–41. [Google Scholar] [CrossRef]
  23. Mollah, M.Y.; Morkovsky, P.; Gomes, J.A.; Kesmez, M.; Parga, J.; Cocke, D.L. Fundamentals, present and future perspectives of electrocoagulation. J. Hazard. Mater. 2004, 114, 199–210. [Google Scholar] [CrossRef]
  24. Tanaka, Y. Water dissociation in ion-exchange membrane electrodialysis. J. Membr. Sci. 2002, 203, 227–244. [Google Scholar] [CrossRef]
  25. Casqueira, R.; Torem, M.; Kohler, H. The removal of zinc from liquid streams by electroflotation. Miner. Eng. 2006, 19, 1388–1392. [Google Scholar] [CrossRef]
  26. Pyo, M.; Moon, I. Indirect Electrochemical Oxidation of Phenol by Ce4+, Controlling Surface Insulation of Au Electrode. Bull. Korean Chem. Soc. 2005, 26, 899–902. [Google Scholar]
  27. Chua, H. Bio-accumulation of environmental residues of rare earth elements in aquatic flora Eichhornia crassipes (Mart.) Solms in Guangdong Province of China. Sci. Total Environ. 1998, 214, 79–85. [Google Scholar] [CrossRef]
  28. Maine, M.A.; Duarte, M.V.; Suñé, N.L. Cadmium uptake by floating macrophytes. Water Res. 2001, 35, 2629–2634. [Google Scholar] [CrossRef]
  29. Ghurye, G.L.; Clifford, D.A.; Tripp, A.R. Combined arsenic and nitrate removal by ion exchange. J. Am. Water Work. Assoc. 1999, 91, 85–96. [Google Scholar] [CrossRef]
  30. Seifert, W.; Liu, X.; Samuelson, L. Influence of arsenic adsorption layers on heterointerfaces in GaInAs/InP quantum well structures. Appl. Phys. Lett. 1993, 62, 949–951. [Google Scholar] [CrossRef]
  31. Hering, J.G.; Chen, P.-Y.; Wilkie, J.A.; Elimelech, M.; Liang, S. Arsenic removal by ferric chloride. J. Am. Water Work. Assoc. 1996, 88, 155–167. [Google Scholar] [CrossRef]
  32. Lorenzen, L.; van Deventer, J.; Landi, W. Factors affecting the mechanism of the adsorption of arsenic species on activated carbon. Miner. Eng. 1995, 8, 557–569. [Google Scholar] [CrossRef]
  33. Ghurye, G.; Clifford, D.; Tripp, A. Iron coagulation and direct microfiltration to remove arsenic from groundwater. J. Am. Water Work. Assoc. 2004, 96, 143–152. [Google Scholar] [CrossRef]
  34. Xiong, T.-T.; Leveque, T.; Austruy, A.; Goix, S.; Schreck, E.; Dappe, V.; Sobanska, S.; Foucault, Y.; Dumat, C. Foliar uptake and metal(loid) bioaccessibility in vegetables exposed to particulate matter. Environ. Geochem. Health 2014, 36, 897–909. [Google Scholar] [CrossRef] [PubMed]
  35. Igwe, J.C.; Abia, A.A.; Okenwa, S.I.; Gbaruko, B.C.; Nwokennaya, E.C. Detoxification of Hg2+, As3+ and Pb2+ metal ions from wastewater by biosorption using modified and unmodified coconut fiber: pH, temperature and particle size effects. Am. Chem. Soc. 2005, 230, U1878–U1879. [Google Scholar]
  36. Marín-Rangel, V.M.; Cortés-Martínez, R.; Villanueva, R.A.C.; Garnica-Romo, M.G.; Martínez-Flores, H.E. As (V) Biosorption in an Aqueous Solution Using Chemically Treated Lemon (Citrus aurantifolia Swingle) Residues. J. Food Sci. 2012, 77, T10–T14. [Google Scholar] [CrossRef] [PubMed]
  37. Maiti, A.; DasGupta, S.; Basu, J.K.; De, S. Batch and Column Study: Adsorption of Arsenate Using Untreated Laterite as Adsorbent. Ind. Eng. Chem. Res. 2008, 47, 1620–1629. [Google Scholar] [CrossRef]
  38. Yazdani, M.R.; Tuutijärvi, T.; Bhatnagar, A.; Vahala, R. Adsorptive removal of arsenic(V) from aqueous phase by feldspars: Kinetics, mechanism, and thermodynamic aspects of adsorption. J. Mol. Liq. 2016, 214, 149–156. [Google Scholar] [CrossRef]
  39. Garg, U.; Kaur, M.; Jawa, G.; Sud, D.; Garg, V. Removal of cadmium (II) from aqueous solutions by adsorption on agricultural waste biomass. J. Hazard. Mater. 2008, 154, 1149–1157. [Google Scholar] [CrossRef]
  40. Bailey, S.E.; Olin, T.J.; Bricka, R.; Adrian, D. A review of potentially low-cost sorbents for heavy metals. Water Res. 1999, 33, 2469–2479. [Google Scholar] [CrossRef]
  41. Hashem, A.; Abdel-Halim, E.S.; El-Tahlawy, K.F.; Hebeish, A. Enhancement of the Adsorption of Co(II) and Ni(II) Ions onto Peanut Hulls through Esterification Using Citric Acid. Adsorpt. Sci. Technol. 2005, 23, 367–380. [Google Scholar] [CrossRef]
  42. Zewail, T.M.; El-Garf, S.A.M. Preparation of agriculture residue based adsorbents for heavy metal removal. Desalin. Water Treat. 2010, 22, 363–370. [Google Scholar] [CrossRef] [Green Version]
  43. Ebrahimi, R.; Maleki, A.; Shahmoradi, B.; Daraei, H.; Mahvi, A.H.; Barati, A.H.; Eslami, A. Elimination of arsenic contamination from water using chemically modified wheat straw. Desalin. Water Treat. 2013, 51, 2306–2316. [Google Scholar] [CrossRef]
  44. Tian, Y.; Wu, M.; Lin, X.; Huang, P.; Huang, Y. Synthesis of magnetic wheat straw for arsenic adsorption. J. Hazard. Mater. 2011, 193, 10–16. [Google Scholar] [CrossRef]
  45. Cheraghi, M.; Lorestani, B.; Merrikhpour, H.; Mosaed, H.P. Assessment efficiency of tea wastes in arsenic removal from aqueous solution. Desalin. Water Treat. 2013, 52, 7235–7240. [Google Scholar] [CrossRef]
  46. Kamala, C.; Chu, K.; Chary, N.; Pandey, P.K.; Ramesh, S.; Sastry, A.; Sekhar, K. Removal of arsenic(III) from aqueous solutions using fresh and immobilized plant biomass. Water Res. 2005, 39, 2815–2826. [Google Scholar] [CrossRef]
  47. Nashine, A.; Tembhurkar, A. Equilibrium, kinetic and thermodynamic studies for adsorption of As(III) on coconut (Cocos nucifera L.) fiber. J. Environ. Chem. Eng. 2016, 4, 3267–3273. [Google Scholar] [CrossRef]
  48. Nashine, A.; Tembhurkar, A. Iron oxide impregnated sugarcane bagasse waste material as sorbent for As(III) removal from water: Kinetic, equilibrium and thermodynamic studies. J. Water Supply Res. Technol. 2016, 65, 645–652. [Google Scholar] [CrossRef]
  49. Urik, M.; Littera, P.; Kolen, M. Removal of arsenic (V) from aqueous solutions using chemically modified sawdust of spruce (Picea abies): Kinetics and isotherm studies. Int. J. Environ. Sci. Technol. 2009, 6, 451–456. [Google Scholar] [CrossRef] [Green Version]
  50. Khaskheli, M.I.; Memon, S.Q.; Siyal, A.N.; Khuhawar, M.Y. Use of Orange Peel Waste for Arsenic Remediation of Drinking Water. Waste Biomass-Valoriz. 2011, 2, 423–433. [Google Scholar] [CrossRef]
  51. Kumari, P.; Sharma, P.; Srivastava, S.; Srivastava, M. Biosorption studies on shelled Moringa oleifera Lamarck seed powder: Removal and recovery of arsenic from aqueous system. Int. J. Miner. Process. 2006, 78, 131–139. [Google Scholar] [CrossRef]
  52. Kamsonlian, S.; Suresh, S.; Ramanaiah, V.; Majumder, C.B.; Chand, S.; Kumar, A. Biosorptive behaviour of mango leaf powder and rice husk for arsenic(III) from aqueous solutions. Int. J. Environ. Sci. Technol. 2012, 9, 565–578. [Google Scholar] [CrossRef] [Green Version]
  53. Kamsonlian, S.; Suresh, S.; Majumder, C.; Chand, S. Biosorption of Arsenic by Mosambi (Citrus limetta) Peel: Equilibrium, Kinetics, Thermodynamics and Desorption Study. Asian J. Chem. 2013, 25, 2409–2417. [Google Scholar] [CrossRef]
  54. Ouédraogo, I.W.K.; Pehlivan, E.; Tran, H.T.; Bonzi-Coulibaly, Y.L.; Zachmann, D.; Bahadir, M. Synthesis of iron oxyhydroxide-coated rice straw (IOC-RS) and its application in arsenic(V) removal from water. J. Water Health 2015, 13, 726–736. [Google Scholar] [CrossRef]
  55. Pehlivan, E.; Tran, T.; Ouédraogo, W.; Schmidt, C.; Zachmann, D.; Bahadir, M. Removal of As(V) from aqueous solutions by iron coated rice husk. Fuel Process. Technol. 2013, 106, 511–517. [Google Scholar] [CrossRef]
  56. Montero, J.I.Z.; Monteiro, A.S.; Gontijo, E.S.; Bueno, C.C.; de Moraes, M.A.; Rosa, A.H. High efficiency removal of As(III) from waters using a new and friendly adsorbent based on sugarcane bagasse and corncob husk Fe-coated biochars. Ecotoxicol. Environ. Saf. 2018, 162, 616–624. [Google Scholar] [CrossRef] [Green Version]
  57. Anirudhan, T.; Unnithan, M.R. Arsenic(V) removal from aqueous solutions using an anion exchanger derived from coconut coir pith and its recovery. Chemosphere 2007, 66, 60–66. [Google Scholar] [CrossRef]
  58. Haldhar, D.; Sahoo, S.; Mishra, P. Adsorption of As (III) From Aqueous Solution by Groundnut Shell. Indian J. Appl. Res. 2014, 4, 2249–2255. [Google Scholar]
  59. Guo, X.; Chen, F. Removal of Arsenic by Bead Cellulose Loaded with Iron Oxyhydroxide from Groundwater. Environ. Sci. Technol. 2005, 39, 6808–6818. [Google Scholar] [CrossRef]
  60. Setyono, D.; Valiyayeettil, S. Chemically Modified Sawdust as Renewable Adsorbent for Arsenic Removal from Water. ACS Sustain. Chem. Eng. 2014, 2, 2722–2729. [Google Scholar] [CrossRef]
  61. Pillai, P.; Kakadiya, N.; Timaniya, Z.; Dharaskar, S.; Sillanpaa, M. Removal of arsenic using iron oxide amended with rice husk nanoparticles from aqueous solution. Mater. Today Proc. 2020, 28, 830–835. [Google Scholar] [CrossRef]
  62. Sarmah, S.; Saikia, J.; Phukan, A.; Goswamee, R.L. Adsorption of As(V) from Water over a Hydroxyl-Alumina Modified Paddy Husk Ash Surface and Its Sludge Immobilization. Water Air Soil Pollut. 2019, 230, 32. [Google Scholar] [CrossRef]
  63. Nasseri, S.; Heidari, M. Evaluation and comparison of aluminum-coated pumice and zeolite in arsenic removal from water resources. Iran. J. Environ. Health Sci. Eng. 2012, 9, 38. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Gupta, V.; Saini, V.; Jain, N. Adsorption of As(III) from aqueous solutions by iron oxide-coated sand. J. Colloid Interface Sci. 2005, 288, 55–60. [Google Scholar] [CrossRef] [PubMed]
  65. Aredes, S.; Klein, B.; Pawlik, M. The removal of arsenic from water using natural iron oxide minerals. J. Clean. Prod. 2012, 29–30, 208–213. [Google Scholar] [CrossRef]
  66. Jeon, C.-S.; Baek, K.; Park, J.-K.; Oh, Y.-K.; Lee, S.-D. Adsorption characteristics of As(V) on iron-coated zeolite. J. Hazard. Mater. 2009, 163, 804–808. [Google Scholar] [CrossRef]
  67. Chen, W.; Parette, R.; Zou, J.; Cannon, F.S.; Dempsey, B.A. Arsenic removal by iron-modified activated carbon. Water Res. 2007, 41, 1851–1858. [Google Scholar] [CrossRef]
  68. Khalid, N.; Ahmad, S.; Toheed, A.; Ahmed, J. Immobilization of Arsenic on Rice Husk. Adsorpt. Sci. Technol. 1998, 16, 655–666. [Google Scholar] [CrossRef]
  69. Manning, B.A.; Goldberg, S. Modeling Competitive Adsorption of Arsenate with Phosphate and Molybdate on Oxide Minerals. Soil Sci. Soc. Am. J. 1996, 60, 121–131. [Google Scholar] [CrossRef] [Green Version]
  70. Amin, M.N.; Kaneco, S.; Kitagawa, C.; Begum, A.; Katsumata, H.; Suzuki, T.; Ohta, K. Removal of arsenic in aqueous solutions by adsorption onto waste rice husk. Ind. Eng. Chem. Res. 2006, 45, 8105–8110. [Google Scholar] [CrossRef]
  71. Doušová, B.; Fuitová, L.; Grygar, T.M.; Machovič, V.; Koloušek, D.; Herzogová, L.; Lhotka, M. Modified aluminosilicates as low-cost sorbents of As(III) from anoxic groundwater. J. Hazard. Mater. 2009, 165, 134–140. [Google Scholar] [CrossRef]
  72. Luengo, C.; Puccia, V.; Avena, M. Arsenate adsorption and desorption kinetics on a Fe(III)-modified montmorillonite. J. Hazard. Mater. 2011, 186, 1713–1719. [Google Scholar] [CrossRef]
  73. Katsoyiannis, I.A.; Zouboulis, A. Removal of arsenic from contaminated water sources by sorption onto iron-oxide-coated polymeric materials. Water Res. 2002, 36, 5141–5155. [Google Scholar] [CrossRef]
  74. Poudel, B.R.; Aryal, R.L.; Bhattarai, S.; Koirala, A.R.; Gautam, S.K.; Ghimire, K.N.; Pant, B.; Park, M.; Paudyal, H.; Pokhrel, M.R. Agro-Waste Derived Biomass Impregnated with TiO2 as a Potential Adsorbent for Removal of As(III) from Water. Catalysts 2020, 10, 1125. [Google Scholar] [CrossRef]
  75. Sorg, T.J.; Wang, L.L. Arsenic removal from drunking water by activated alumina and anion exchange treatment. In Proceedings of the Small Drinking Water and Wastewater Systems, Phoenix, AZ, USA, 12 January 2000; pp. 282–291. [Google Scholar]
  76. Maiti, A.; Dasgupta, S.; Basu, J.K.; De, S. Adsorption of arsenite using natural laterite as adsorbent. Sep. Purif. Technol. 2007, 55, 350–359. [Google Scholar] [CrossRef]
  77. Twidwell, L.G.; Robins, R.G.; Hohn, J.W. The removal of arsenic from aqueous solution by coprecipitation with iron (III). In Proceedings of the Arsenic Metallurgy, Fundamentals and Applications, TMS Annual Meeting, San Francisco, CA, USA, 13–17 February 2005; pp. 3–24. [Google Scholar]
  78. Roberts, L.C.; Hug, S.J.; Ruettimann, T.; Billah, M.; Khan, A.W.; Rahman, M.T. Arsenic Removal with Iron(II) and Iron(III) in Waters with High Silicate and Phosphate Concentrations. Environ. Sci. Technol. 2004, 38, 307–315. [Google Scholar] [CrossRef]
  79. Maji, S.; Kao, Y.-H.; Liu, C.-W. Arsenic removal from real arsenic-bearing groundwater by adsorption on iron-oxide-coated natural rock (IOCNR). Desalination 2011, 280, 72–79. [Google Scholar] [CrossRef]
  80. Joshi, A.; Chaudhuri, M. Removal of Arsenic from Ground Water by Iron Oxide-Coated Sand. J. Environ. Eng. 1996, 122, 769–771. [Google Scholar] [CrossRef]
  81. Wu, K.; Liu, T.; Peng, J.M. Adsorption Behaviors of Arsenic(V) onto Fe-Based Backwashing Sludge Produced from Fe(II)-Removal Plants. Appl. Mech. Mater. 2013, 295–298, 1321–1326. [Google Scholar] [CrossRef]
  82. Asif, Z.; Chen, Z. Removal of arsenic from drinking water using rice husk. Appl. Water Sci. 2017, 7, 1449–1458. [Google Scholar] [CrossRef] [Green Version]
  83. Gupta, A.; Sankararamakrishnan, N. Column studies on the evaluation of novel spacer granules for the removal of arsenite and arsenate from contaminated water. Bioresour. Technol. 2010, 101, 2173–2179. [Google Scholar] [CrossRef]
  84. Singh, T.S.; Pant, K. Experimental and modelling studies on fixed bed adsorption of As(III) ions from aqueous solution. Sep. Purif. Technol. 2006, 48, 288–296. [Google Scholar] [CrossRef]
  85. Bhakat, P.; Gupta, A.; Ayoob, S. Feasibility analysis of As(III) removal in a continuous flow fixed bed system by modified calcined bauxite (MCB). J. Hazard. Mater. 2007, 139, 286–292. [Google Scholar] [CrossRef]
  86. Deliyanni, E.A.; Bakoyannakis, D.N.; Zouboulis, A.; Peleka, E. Removal of Arsenic and Cadmium by Akaganeite Fixed-Beds. Sep. Sci. Technol. 2003, 38, 3967–3981. [Google Scholar] [CrossRef]
  87. Wu, C.; Huang, L.; Xue, S.-G.; Huang, Y.-Y.; Hartley, W.; Cui, M.-Q.; Wong, M.H. Arsenic sorption by red mud-modified biochar produced from rice straw. Environ. Sci. Pollut. Res. 2017, 24, 18168–18178. [Google Scholar] [CrossRef] [PubMed]
  88. Shafique, U.; Ijaz, A.; Salman, M.; Zaman, W.U.; Jamil, N.; Rehman, R.; Javaid, A. Removal of arsenic from water using pine leaves. J. Taiwan Inst. Chem. Eng. 2012, 43, 256–263. [Google Scholar] [CrossRef]
  89. Liu, C.; Wang, B.; Deng, Y.; Cui, B.; Wang, J.; Chen, W.; He, S.-Y. Performance of a New Magnetic Chitosan Nanoparticle to Remove Arsenic and Its Separation from Water. J. Nanomater. 2015, 2015, 1–9. [Google Scholar] [CrossRef] [Green Version]
  90. Wang, S.; Gao, B.; Li, Y.; Wan, Y.; Creamer, A.E. Sorption of arsenate onto magnetic iron–manganese (Fe–Mn) biochar composites. RSC Adv. 2015, 5, 67971–67978. [Google Scholar] [CrossRef]
  91. Pandey, P.K.; Choubey, S.; Verma, Y.; Pandey, M.; Chandrashekhar, K. Biosorptive removal of arsenic from drinking water. Bioresour. Technol. 2009, 100, 634–637. [Google Scholar] [CrossRef]
  92. Fan, T.; Liu, Y.; Feng, B.; Zeng, G.; Yang, C.; Zhou, M.; Zhou, H.; Tan, Z.; Wang, X. Biosorption of cadmium(II), zinc(II) and lead(II) by Penicillium simplicissimum: Isotherms, kinetics and thermodynamics. J. Hazard. Mater. 2008, 160, 655–661. [Google Scholar] [CrossRef]
  93. Podder, M.; Majumder, C. Predictive approach for simultaneous biosorption and bioaccumulation of arsenic by Corynebacterium glutamicum MTCC 2745 biofilm supported on NL/MnFe2O4 composite. J. Water Process. Eng. 2016, 11, 8–31. [Google Scholar] [CrossRef]
  94. Podder, M.; Majumder, C. Study of the kinetics of arsenic removal from wastewater using Bacillus arsenicus biofilms supported on a Neem leaves/MnFe2O4 composite. Ecol. Eng. 2016, 88, 195–216. [Google Scholar] [CrossRef]
  95. Ansone-Bertina, L.; Klavins, M.; Viksna, A. Arsenic removal using natural biomaterial-based sorbents. Environ. Geochem. Health 2013, 35, 633–642. [Google Scholar] [CrossRef]
  96. Ansone-Bertina, L.; Klavins, M.; Eglite, L. Use of peat-based sorbents for removal of arsenic compounds. Open Chem. 2013, 11, 988–1000. [Google Scholar] [CrossRef]
  97. Verma, L.; Siddique, M.A.; Singh, J.; Bharagava, R.N. As(III) and As(V) removal by using iron impregnated biosorbents derived from waste biomass of Citrus limmeta (peel and pulp) from the aqueous solution and ground water. J. Environ. Manag. 2019, 250, 109452. [Google Scholar] [CrossRef]
  98. Camacho, L.M.; Parra, R.R.; Deng, S. Arsenic removal from groundwater by MnO2-modified natural clinoptilolite zeolite: Effects of pH and initial feed concentration. J. Hazard. Mater. 2011, 189, 286–293. [Google Scholar] [CrossRef]
  99. Wang, S.; Gao, B.; Zimmerman, A.; Li, Y.; Ma, L.; Harris, W.G.; Migliaccio, K. Removal of arsenic by magnetic biochar prepared from pinewood and natural hematite. Bioresour. Technol. 2015, 175, 391–395. [Google Scholar] [CrossRef]
  100. Ren, X.; Zhang, Z.; Luo, H.; Hu, B.; Dang, Z.; Yang, C.; Li, L. Adsorption of arsenic on modified montmorillonite. Appl. Clay Sci. 2014, 97–98, 17–23. [Google Scholar] [CrossRef]
  101. Chen, Y.-N.; Chai, L.-Y.; Shu, Y.-D. Study of arsenic(V) adsorption on bone char from aqueous solution. J. Hazard. Mater. 2008, 160, 168–172. [Google Scholar] [CrossRef]
  102. Fufa, F.; Alemayehu, E.; Lennartz, B. Sorptive removal of arsenate using termite mound. J. Environ. Manag. 2014, 132, 188–196. [Google Scholar] [CrossRef]
  103. Kundu, S.; Gupta, A. Analysis and modeling of fixed bed column operations on As(V) removal by adsorption onto iron oxide-coated cement (IOCC). J. Colloid Interface Sci. 2005, 290, 52–60. [Google Scholar] [CrossRef]
  104. Danish, M.I.; Qazi, I.A.; Zeb, A.; Habib, A.; Awan, M.A.; Khan, Z. Arsenic Removal from Aqueous Solution Using Pure and Metal-Doped Titania Nanoparticles Coated on Glass Beads: Adsorption and Column Studies. J. Nanomater. 2013, 2013, 1–17. [Google Scholar] [CrossRef]
  105. Maji, S.K.; Pal, A.; Pal, T. Arsenic removal from aqueous solutions by adsorption on laterite soil. J. Environ. Sci. Health Part A 2007, 42, 453–462. [Google Scholar] [CrossRef] [PubMed]
  106. Maji, S.K.; Pal, A.; Pal, T.; Adak, A. Sorption kinetics of arsenic on laterite soil in aqueous medium. J. Environ. Sci. Health Part A 2007, 42, 989–996. [Google Scholar] [CrossRef] [PubMed]
  107. Kundu, S.; Kavalakatt, S.S.; Pal, A.; Ghosh, S.K.; Mandal, M.; Pal, T. Removal of arsenic using hardened paste of Portland cement: Batch adsorption and column study. Water Res. 2004, 38, 3780–3790. [Google Scholar] [CrossRef] [PubMed]
  108. Altundoğan, H.S.; Altundoğan, S.; Tümen, F.; Bildik, M. Arsenic removal from aqueous solutions by adsorption on red mud. Waste Manag. 2000, 20, 761–767. [Google Scholar] [CrossRef]
  109. Zhang, W.; Singh, P.; Paling, E.; Delides, S. Arsenic removal from contaminated water by natural iron ores. Miner. Eng. 2004, 17, 517–524. [Google Scholar] [CrossRef]
  110. Roy, P.; Mondal, N.K.; Bhattacharya, S.; Das, B.; Das, K. Removal of arsenic(III) and arsenic(V) on chemically modified low-cost adsorbent: Batch and column operations. Appl. Water Sci. 2013, 3, 293–309. [Google Scholar] [CrossRef] [Green Version]
  111. Korpayev, S.; Kavaklı, C.; Tilki, S.; Kavaklı, P.A. Novel cotton fabric adsorbent for efficient As(V) adsorption. Environ. Sci. Pollut. Res. 2018, 25, 34610–34622. [Google Scholar] [CrossRef]
  112. Alam Khan, T.; Chaudhry, S.A.; Ali, I. Thermodynamic and kinetic studies of As(V) removal from water by zirconium oxide-coated marine sand. Environ. Sci. Pollut. Res. 2013, 20, 5425–5440. [Google Scholar] [CrossRef]
  113. Hansen, H.K.; Ribeiro, A.; Mateus, E. Biosorption of arsenic(V) with Lessonia nigrescens. Miner. Eng. 2006, 19, 486–490. [Google Scholar] [CrossRef]
  114. Raichur, A.; Panvekar, V. Removal of As(V) by adsorption onto mixed rare earth oxides. Sep. Sci. Technol. 2002, 37, 1095–1108. [Google Scholar] [CrossRef]
  115. Dos Santos, H.H.; Demarchi, C.A.; Rodrigues, C.A.; Greneche, J.M.; Nedelko, N.; Slawska-Waniewska, A. Adsorption of As(III) on chitosan-Fe-crosslinked complex (Ch-Fe). Chemosphere 2011, 82, 278–283. [Google Scholar] [CrossRef]
  116. Manju, G.N.; Raji, C.; Anirudhan, T.S. Evaluation of coconut husk carbon for the removal of arsenic from water. Water Res. 1998, 32, 3062–3070. [Google Scholar] [CrossRef]
  117. Polowczyk, I.; Bastrzyk, A.; Ulatowska, J.; Szczałba, E.; Koźlecki, T.; Sadowski, Z. Influence of pH on arsenic(III) removal by fly ash. Sep. Sci. Technol. 2016, 51, 2612–2619. [Google Scholar] [CrossRef]
  118. Giménez, J.; Martinez, M.; De Pablo, J.; Rovira, M.; Duro, L. Arsenic sorption onto natural hematite, magnetite, and goethite. J. Hazard. Mater. 2007, 141, 575–580. [Google Scholar] [CrossRef]
  119. Ben Issa, N.; Rajaković-Ognjanović, V.N.; Marinković, A.; Rajaković, L.V. Separation and determination of arsenic species in water by selective exchange and hybrid resins. Anal. Chim. Acta 2011, 706, 191–198. [Google Scholar] [CrossRef]
  120. Chuang, C.; Fan, M.; Xu, M.; Brown, R.; Sung, S.; Saha, B. Adsorption of arsenic(V) by activated carbon prepared from oat hulls. Chemosphere 2005, 61, 478–483. [Google Scholar] [CrossRef]
  121. Swedlund, P.J.; Webster, J.G. Adsorption and polymerisation of silicic acid on ferrihydrite, and its effect on arsenic adsorption. Water Res. 1999, 33, 3413–3422. [Google Scholar] [CrossRef]
  122. Niazi, N.K.; Bibi, I.; Shahid, M.; Ok, Y.S.; Burton, E.; Wang, H.; Shaheen, S.; Rinklebe, J.; Lüttge, A. Arsenic removal by perilla leaf biochar in aqueous solutions and groundwater: An integrated spectroscopic and microscopic examination. Environ. Pollut. 2018, 232, 31–41. [Google Scholar] [CrossRef]
  123. Wang, S.; Gao, B.; Li, Y. Enhanced arsenic removal by biochar modified with nickel (Ni) and manganese (Mn) oxyhydroxides. J. Ind. Eng. Chem. 2016, 37, 361–365. [Google Scholar] [CrossRef] [Green Version]
  124. Wang, S.; Gao, B.; Li, Y.; Creamer, A.E.; He, F. Adsorptive removal of arsenate from aqueous solutions by biochar supported zero-valent iron nanocomposite: Batch and continuous flow tests. J. Hazard. Mater. 2017, 322, 172–181. [Google Scholar] [CrossRef]
  125. He, R.; Peng, Z.; Lyu, H.; Huang, H.; Nan, Q.; Tang, J. Synthesis and characterization of an iron-impregnated biochar for aqueous arsenic removal. Sci. Total Environ. 2018, 612, 1177–1186. [Google Scholar] [CrossRef]
  126. Xu, Y.-H.; Nakajima, T.; Ohki, A. Adsorption and removal of arsenic(V) from drinking water by aluminum-loaded Shirasu-zeolite. J. Hazard. Mater. 2002, 92, 275–287. [Google Scholar] [CrossRef]
  127. Xiong, Y.; Tong, Q.; Shan, W.; Xing, Z.; Wang, Y.; Wen, S.; Lou, Z. Arsenic transformation and adsorption by iron hydroxide/manganese dioxide doped straw activated carbon. Appl. Surf. Sci. 2017, 416, 618–627. [Google Scholar] [CrossRef]
  128. Zhang, F.; Wang, X.; Xionghui, J.; Ma, L. Efficient arsenate removal by magnetite-modified water hyacinth biochar. Environ. Pollut. 2016, 216, 575–583. [Google Scholar] [CrossRef]
  129. Tuzen, M.; Sari, A.; Mendil, D.; Uluozlu, O.D.; Soylak, M.; Dogan, M. Characterization of biosorption process of As(III) on green algae Ulothrix cylindricum. J. Hazard. Mater. 2009, 165, 566–572. [Google Scholar] [CrossRef]
  130. Wang, S.; Gao, B.; Li, Y.; Zimmerman, A.R.; Cao, X. Sorption of arsenic onto Ni/Fe layered double hydroxide (LDH)-biochar composites. RSC Adv. 2016, 6, 17792–17799. [Google Scholar] [CrossRef]
  131. Abid, M.; Niazi, N.K.; Bibi, I.; Farooqi, A.; Ok, Y.S.; Kunhikrishnan, A.; Ali, F.; Ali, S.; Igalavithana, A.D.; Arshad, M. Arsenic(V) biosorption by charred orange peel in aqueous environments. Int. J. Phytoremediat. 2016, 18, 442–449. [Google Scholar] [CrossRef]
  132. Shakoor, M.B.; Niazi, N.K.; Bibi, I.; Shahid, M.; Sharif, F.; Bashir, S.; Shaheen, S.M.; Wang, H.; Tsang, D.C.W.; Ok, Y.S.; et al. Arsenic removal by natural and chemically modified watermelon rind in aqueous solutions and groundwater. Sci. Total Environ. 2018, 645, 1444–1455. [Google Scholar] [CrossRef] [PubMed]
  133. Gupta, A.; Vidyarthi, S.R.; Sankararamakrishnan, N. Concurrent removal of As(III) and As(V) using green low cost functionalized biosorbent—Saccharum officinarum bagasse. J. Environ. Chem. Eng. 2015, 3, 113–121. [Google Scholar] [CrossRef]
  134. Maiti, A.; Basu, J.K.; De, S. Removal of arsenic from synthetic and natural groundwater using acid-activated laterite. Environ. Prog. Sustain. Energy 2010, 29, 457–470. [Google Scholar] [CrossRef]
  135. Baig, S.A.; Zhu, J.; Muhammad, N.; Sheng, T.; Xu, X. Effect of synthesis methods on magnetic Kans grass biochar for enhanced As(III, V) adsorption from aqueous solutions. Biomass Bioenergy 2014, 71, 299–310. [Google Scholar] [CrossRef]
  136. Bayramoglu, G.; Altintas, B.; Arica, M.Y. Adsorption kinetics and thermodynamic parameters of cationic dyes from aqueous solutions by using a new strong cation-exchange resin. Chem. Eng. J. 2009, 152, 339–346. [Google Scholar] [CrossRef]
  137. Liu, C.-C.; Kuang-Wang, M.; Li, Y.-S. Removal of Nickel from Aqueous Solution Using Wine Processing Waste Sludge. Ind. Eng. Chem. Res. 2005, 44, 1438–1445. [Google Scholar] [CrossRef]
  138. Cope, C.O.; Webster, D.S.; Sabatini, D.A. Arsenate adsorption onto iron oxide amended rice husk char. Sci. Total Environ. 2014, 488–489, 554–561. [Google Scholar] [CrossRef]
  139. Zhou, Z.; Liu, Y.-G.; Liu, S.-B.; Liu, H.-Y.; Zeng, G.-M.; Tan, X.; Yang, C.-P.; Ding, Y.; Yan, Z.-L.; Cai, X. Sorption performance and mechanisms of arsenic(V) removal by magnetic gelatin-modified biochar. Chem. Eng. J. 2017, 314, 223–231. [Google Scholar] [CrossRef]
  140. Pokhrel, D.; Viraraghavan, T. Arsenic removal from an aqueous solution by modified A. niger biomass: Batch kinetic and isotherm studies. J. Hazard. Mater. 2008, 150, 818–825. [Google Scholar] [CrossRef]
  141. Chammui, Y.; Sooksamiti, P.; Naksata, W.; Thiansem, S.; Arqueropanyo, O.-A. Removal of arsenic from aqueous solution by adsorption on Leonardite. Chem. Eng. J. 2014, 240, 202–210. [Google Scholar] [CrossRef]
  142. Baig, S.A.; Sheng, T.; Sun, C.; Xue, X.; Tan, L.; Xu, X. Arsenic Removal from Aqueous Solutions Using Fe3O4-HBC Composite: Effect of Calcination on Adsorbents Performance. PLoS ONE 2014, 9, e100704. [Google Scholar] [CrossRef]
  143. Lin, L.; Qiu, W.; Wang, D.; Huang, Q.; Song, Z.; Chau, H.W. Arsenic removal in aqueous solution by a novel Fe-Mn modified biochar composite: Characterization and mechanism. Ecotoxicol. Environ. Saf. 2017, 144, 514–521. [Google Scholar] [CrossRef]
  144. Hu, X.; Ding, Z.; Zimmerman, A.; Wang, S.; Gao, B. Batch and column sorption of arsenic onto iron-impregnated biochar synthesized through hydrolysis. Water Res. 2015, 68, 206–216. [Google Scholar] [CrossRef]
  145. Guo, H.; Stüben, D.; Berner, Z.A. Removal of arsenic from aqueous solution by natural siderite and hematite. Appl. Geochem. 2007, 22, 1039–1051. [Google Scholar] [CrossRef]
  146. Turk, T. Removal of Dissolved Arsenic by Pyrite Ash Waste. Mine Water Environ. 2017, 36, 255–263. [Google Scholar] [CrossRef]
  147. Samsuri, W.; Sadegh-Zadeh, F.; Seh-Bardan, B.J. Adsorption of As(III) and As(V) by Fe coated biochars and biochars produced from empty fruit bunch and rice husk. J. Environ. Chem. Eng. 2013, 1, 981–988. [Google Scholar] [CrossRef]
  148. Cedillo, M.J.J.; Olguín, M.; Fall, C.; Colin-Cruz, A. Adsorption capacity of iron- or iron–manganese-modified zeolite-rich tuffs for As(III) and As(V) water pollutants. Appl. Clay Sci. 2011, 54, 206–216. [Google Scholar] [CrossRef]
  149. Baskan, M.B.; Pala, A. Removal of arsenic from drinking water using modified natural zeolite. Desalination 2011, 281, 396–403. [Google Scholar] [CrossRef]
  150. Ho, Y.-S.; Ofomaja, A. Kinetics and thermodynamics of lead ion sorption on palm kernel fibre from aqueous solution. Process. Biochem. 2005, 40, 3455–3461. [Google Scholar] [CrossRef]
  151. Ho, Y.-S.; Ofomaja, A. Effects of calcium competition on lead sorption by palm kernel fibre. J. Hazard. Mater. 2005, 120, 157–162. [Google Scholar] [CrossRef] [PubMed]
  152. Zhang, M.; Gao, B.; Varnoosfaderani, S.; Hebard, A.; Yao, Y.; Inyang, M. Preparation and characterization of a novel magnetic biochar for arsenic removal. Bioresour. Technol. 2013, 130, 457–462. [Google Scholar] [CrossRef] [PubMed]
  153. Kong, S.; Wang, Y.; Hu, Q.; Olusegun, A.K. Magnetic nanoscale Fe–Mn binary oxides loaded zeolite for arsenic removal from synthetic groundwater. Colloids Surf. A Physicochem. Eng. Asp. 2014, 457, 220–227. [Google Scholar] [CrossRef]
  154. Ho, Y.S. Citation review of Lagergren kinetic rate equation on adsorption reactions. Scientometrics 2004, 59, 171–177. [Google Scholar]
  155. Ho, Y.S.; McKay, G. Pseudo-second order model for sorption processes. Process. Biochem. 1999, 34, 451–465. [Google Scholar] [CrossRef]
  156. Jiménez-Cedillo, M.J.; Olguín, M.T.; Fall, C.; Colin-Cruz, A. As(III) and As(V) sorption on iron-modified non-pyrolyzed and pyrolyzed biomass from Petroselinum crispum (parsley). J. Environ. Manag. 2013, 117, 242–252. [Google Scholar] [CrossRef]
  157. Dorraji, M.S.S.; Mirmohseni, A.; Tasselli, F.; Criscuoli, A.; Carraro, M.; Gross, S.; Figoli, A. Preparation, characterization and application of iron (III)-loaded chitosan hollow fiber membranes as a new bio-based As (V) sorbent. J. Polym. Res. 2014, 21, 399. [Google Scholar] [CrossRef]
  158. Xia, D.; Tan, F.; Zhang, C.; Jiang, X.; Chen, Z.; Li, H.; Zheng, Y.; Li, Q.; Wang, Y. ZnCl2-activated biochar from biogas residue facilitates aqueous As(III) removal. Appl. Surf. Sci. 2016, 377, 361–369. [Google Scholar] [CrossRef]
  159. Mohan, D.; Pittman, C.U.; Bricka, M.; Smith, F.; Yancey, B.; Mohammad, J.; Steele, P.H.; Franco, M.A.; Gómez-Serrano, V.; Gong, H. Sorption of arsenic, cadmium, and lead by chars produced from fast pyrolysis of wood and bark during bio-oil production. J. Colloid Interface Sci. 2007, 310, 57–73. [Google Scholar] [CrossRef]
Figure 1. Different types of natural adsorbent As.
Figure 1. Different types of natural adsorbent As.
Coatings 11 01407 g001
Figure 2. Functional groups existing on the surface of natural adsorbents.
Figure 2. Functional groups existing on the surface of natural adsorbents.
Coatings 11 01407 g002
Figure 3. As(III) adsorption as a function of pH variation [116,117].
Figure 3. As(III) adsorption as a function of pH variation [116,117].
Coatings 11 01407 g003
Figure 4. As(V) adsorption as a function of pH variation [123,124].
Figure 4. As(V) adsorption as a function of pH variation [123,124].
Coatings 11 01407 g004
Table 1. Some As natural adsorbents and their adsorption capacities.
Table 1. Some As natural adsorbents and their adsorption capacities.
AdsorbentsTreatmentsIsotherm Model FitpHAdsorption Capacities mg/g
As(III)As(V)
Wheat Straw [43]NaHCO3Langmuir7No data0.097
Wheat Straw [44]FeSO4 and FeCl3Langmuir6No data30.24
Black Tea [45]No TreatmentsFreundlich and Langmuir6No data1.76
G. cambogia fruits rinds [46]IICT TechnologyLangmuir6128.10No data
Coconut Fibers [47]HNO3 and NaOHFreundlich40.12No data
Sugarcane bagasse [48](Fe(NO3)3·9H2O)Langmuir60.6No data
Sawdust [49]Fe(III)Langmuir8No data5.8
Orange Peel [50]No TreatmentsFreundlich7No data4.8
Moringa Lamarck seed [51]No TreatmentsLangmuir81.52.1
Rice Husk [52]No TreatmentsFreundlich7220.1No data
Mango leaf powder [52]No TreatmentsFreundlich7250.07No data
Mosambi Citrus Peel [53]No TreatmentsFreundlich62.123.32
Rice straw [54]Fe(NO3)3Langmuir4No data21.739
Rice husk [55]Fe(III)Langmuir4No data2.47
Corncob husk [56]FeCl3Langmuir650No data
Coconut coir pith (CP) [57]Epichlorohydrin and dimethylamineLangmuir7No data13.57
Groundnut shells [58]No TreatmentsLangmuir80.014No data
Bead Cellulose (Cotton) [59](FeCl3‚6H2O)Langmuir9499.633.2
Sawdust [60]ZrOCl₂·8H₂OLangmuir942912
Rice husk [61]Iron oxideLangmuir6No data82
Paddy Husk Ash [62]AlCl3·6H2OFreundlich4No data0.063
Zeolite stones [63]Al2(SO4)·16H2OFreundlich6No data208
Iron oxide-coated sand [64]Fe(III)Langmuir7.528.57No data
Goethite [65]No treatmentNo data5No data1
Iron-coated zeolite [66]FeCl3Langmuir4No data0.68
Iron-modified activated carbon [67]Fe(NO3)3·9H2OLangmuir8643.651.3
Table 2. Some studies related to initial As concentration and adsorbent dosage impact on As adsorption capacity.
Table 2. Some studies related to initial As concentration and adsorbent dosage impact on As adsorption capacity.
AdsorbentAdsorbent Dosage (g)Solution Quantity (mL)Intial Concentration (mg/L)Removal (%)
As(V)As(III)
Peat-based sorbents [96]0.5408–904No data100–43
Red mud-modified biochar (RM-BC) [87]0.12301–50100–3250–4
Novel magnetic chitosan nanoparticle (MCNP) [89]0.051000.2–50100–65100–60
Hematite Pinewood biochar (HPB) [99]0.05201–5025–2No data
Modified clinoptilolite zeolite (MCZ) [98]12000.005–0.0555–40No data
Iron-modified peat [96]0.540100–270100–70No data
Biomass of Citrus limmeta (PPAC-500) [97] 0.15500.05–2.5100–6293–55.2
Leaves of P. roxburghii powder [88]0.2–250104–75No data
Iron modified montmorillonites [100]0.025–0.3250.00597–10096–100
Corynebacterium glutamicum MTCC 2745 biofilm supported on Neem leaves NL/MnFe2O4 composite [93]0.01–0.11005079–82.572–77
Bacillus arsenicus biofilms supported on a Neem leaves/MnFe2O4 composite [94]0.01–0.11005086–8979–83
M. charantia plants biomass [91]0.05–0.25500.5No data66–88
Bone char [101]0.05–0.45000.530–100No data
Table 3. Studies on some natural adsorbent’s reusability performance.
Table 3. Studies on some natural adsorbent’s reusability performance.
AdsorbentAdsorbent Dosage (g)Intial Concentration (mg/L)RegenetaionAdsorption Capacity (%)Reference
As(V)As(III)
The green alga
(U. cylindricum) biomass [129]
0.11010No data96–93Tuzen M. et al., 2009
Ni/Mn-layered double
hydroxide (LDH) biochar (NMMB) [123]
0.1403100–98No dataWang S. S. et al., 2016a
Ni/Fe layered double
hydroxide (LDH)-biochar (NFMB) [130]
0.1503100–92No dataWang S. S. et al., 2016b
Charred orange peel (COP) [131]0.22003100–90No data Abid M. et al., 2016
Fe-Mn-straw biochars [127]1203No data85–78Xiong Y. et al., 2017
Xanthated water melon rind (X-WMR) [132]0.2As(V)
5
As(III)
4
4100–50100–20Shakoor M. B. et al., 2018
Magnetite-modified water hyacinth
Biochar (MW2501) [128]
0.254100–50No dataZhang F. et al., 2016
20% Iron-impregnated corn straw biochar [125] 0.240387–70No dataHe R. Z. et al., 2018
Modified Saccharum officinarum bagasse (SCB-S) [133]0.050.55100–78 100–86Gupta A. et al., 2015
Table 4. Other ions influence on some natural adsorbents.
Table 4. Other ions influence on some natural adsorbents.
AdsorbentsThermodynamicsKeneticInfluence Ions (mg/L)Removal As (%)
As(III)As(V)
Acid-activated laterite (AAL) [134]EndothermicPseudo-second-orderPO43−0–1089–7095–85
SO42−0–4089–6895–83
Fe3O4-HBC-1000 °C
(Honeycomb Briquette Cinders) [142]
EndothermicPseudo-second-orderPO43−0–9.5100–15100–60
Fe-Mn modified corn stem biochar (FMBC) [143]EndothermicPseudo-second-orderPO43−0–9500100–88No data
Iron oxide amended rice husk char
(950 IOA-RHC) [138]
EndothermicPseudo-second-orderPO43−0–10No data100–65
Fe-impregnated hickory chips biochar [144]No dataNo dataPO43−0–50No data100–15
Magnetic Kans Grass (Saccharum spontaneum) Biochars (MKGB4) [135]No dataPseudo-second-orderPO43−0–95100–20100–80
Leonardite chars [141]No dataNo dataSO42−0–500098–9299–94
Siderite SIO3 [145]No dataNo dataPO43−0–10No data64–28
Hematite HIO1 [145]No dataFirst-orderPO43−0–10No data69–36
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

YEO, K.F.H.; Li, C.; Zhang, H.; Chen, J.; Wang, W.; Dong, Y. Arsenic Removal from Contaminated Water Using Natural Adsorbents: A Review. Coatings 2021, 11, 1407. https://doi.org/10.3390/coatings11111407

AMA Style

YEO KFH, Li C, Zhang H, Chen J, Wang W, Dong Y. Arsenic Removal from Contaminated Water Using Natural Adsorbents: A Review. Coatings. 2021; 11(11):1407. https://doi.org/10.3390/coatings11111407

Chicago/Turabian Style

YEO, Kanfolo Franck Herve, Chaokun Li, Hui Zhang, Jin Chen, Wendong Wang, and Yingying Dong. 2021. "Arsenic Removal from Contaminated Water Using Natural Adsorbents: A Review" Coatings 11, no. 11: 1407. https://doi.org/10.3390/coatings11111407

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