Khalid Z.
Elwakeel
*ab,
Ahmed M.
Elgarahy
c,
Ziya A.
Khan
a,
Muath S.
Almughamisi
a and
Abdullah S.
Al-Bogami
a
aUniversity of Jeddah, College of Science, Department of Chemistry, Jeddah, Saudi Arabia. E-mail: kelwkeel@uj.edu.sa; khalid_elwakeel@sci.psu.edu.eg
bEnvironmental Science Department, Faculty of Science, Port-Said University, Port-Said, Egypt
cZoology Department, Faculty of Science, Port-Said University, Port-Said, Egypt
First published on 1st June 2020
Metal/mineral-incorporating materials have received significant attention over the last decades given the outstanding adsorption behavior towards various pollutants, especially Cr(VI), in aqueous solutions. Here, the pattern of sorption of some pollutants with special focus on Cr(VI) removal over metal/mineral-incorporating materials has been compiled. Furthermore, the key influencing adsorption variables, i.e., pH, concentration at the beginning, contact time, and dosage of sorbent, were discussed while considering different material classifications. Different isothermal and kinetic models were elaborated. Langmuir and Freundlich's models are adopted as the main sorption isotherms, while the pseudo-second-order kinetic model is the most fitted heavy metal ion and Cr(VI) kinetic model in aqueous systems. The results revealed that metal/mineral-incorporating materials are quite effective for heavy metal (especially Cr(VI)) recovery from water and confirmed that these materials are affordable and reliable for contaminated water remediation. Also, several methods are available for the modification of these materials in order to increase their sorption efficiency. However, to establish the use of metal/mineral-incorporating materials for water purification compared with other established methods, more investigations are required to determine the best modification method and investigate the release of metals from these materials during sorption.
In environmental pollution, heavy metals mainly refer to mercury, cadmium, chromium, lead, and arsenic, which have significant biological toxicity. Usually, the natural background concentrations of heavy metals do not reach harmful levels. Poisonous and harmful metals get into the atmosphere, water, and soil and cause serious environmental pollution because of human activities such as heavy metal mining. As major toxicants, heavy metals are found in industrial wastewater and possess severe effects on wastewater biological treatment. Heavy metals’ sources in wastewater treatment plants are from industrial discharge and urban stormwater runoff. Some of the heavy metals, e.g., Cd, Cr, Pb, Hg, and Ag, are priority pollutants. Heavy metal toxicity is caused by the fact that they are soluble metals. The toxicity is managed by various factors such as pH and the type and concentration of complexing agents in wastewater.9 Due to their discharge into the environment, heavy metal concentrations in sediments and suspended solids in water have increased sharply and major water pollution problems have occurred.10 Many heavy metals, especially at high concentrations, are deadly to even the most resistant species of bacteria, algae, and fungi. Therefore, in many cases of highly contaminated industrial wastewater, biological treatment would be impractical or even impossible.11 High concentrations of trace metals can also be found in groundwater near contaminated sources; however, these may pose potential health threats. Some trace constituents are associated with industrial pollution, such as As(V) and Cr(VI).12,13 Chromium is present in the zerovalent state, while in the environment, trivalent and hexavalent chromium species are the most thermodynamically stable. Hexavalent chromium usually occurs as CrO42− or HCrO4−, whereas Cr(III) as Cr(OH)n(3−n)+. Chromium(III) is an essential nutrient while Cr(VI) is strongly carcinogenic.14 It has long been recognized that Cr(III) occurs naturally in most environmental media, while Cr(VI) only naturally in groundwater.
The World Health Organization recommended the total value of chromium in drinking water to be 0.05 mg L−1.15 The European Union has also recommended the same value in its council directive 98/83/EC on drinkable water quality.16 However, the United States Environmental Protection Agency has established 0.1 mg L−1 as the maximum allowed contaminant level for total chromium.17
Chromium(VI) is one of the most toxic heavy metals existing in wastewater. Until now, various adsorbents were broadly tested for the displacement of Cr(VI) from contaminated water, but the results illustrate a low and hindered adsorption capacity. Chromium, as well as its compounds, is usually used in steel production and metallurgy, the chemical industry, leather tanning, protection and decorative coatings, the automotive industry, pigment manufacture, and wood treatment. Due to such an extensive range of use, large amounts of chromium are released into the atmosphere and terrestrial and aquatic environments. As a result of developed industrial activities, huge amounts of chromium-rich waste materials are generated.18 Waste from the steelmaking industry may often be reused in civil engineering, while other chromium-containing waste materials are in general disposed of.19 In nature, under environmental conditions, Cr(III) species prevail, while Cr(VI) compounds are mainly of anthropogenic origin. The essentiality and toxic nature of Cr depend primarily on its chemical forms. Generally, Cr(III) compounds are of lower toxicity compared with Cr(VI). Cr(III) is a crucial micronutrient present in glucose and lipid metabolism. For minimizing the health hazards of Cr(VI) to occupationally exposed workers, the Occupational Safety and Health Administration standardized the limits of exposure to Cr(VI) compounds in the workplace. Chromium(VI) usage is limited by different regulative acts, and measures are also taken to diminish the environmental impacts of disposed of and reused chromium waste materials.20 The soluble Cr(VI) species are readily released into soil solution and ground- and surface water and can be taken up by plants. To prevent harmful effects of Cr(VI) on terrestrial and aquatic environments and safely dispose of or reuse chromium-rich waste materials, reliable determination of Cr(VI) is of paramount importance.21 Accurately determining Cr(VI) is also crucial in the investigations of the oxidation–reduction processes of Cr in soil and the evaluation of the efficiencies of the remediation processes of Cr-contaminated soil. Chromium(VI) exists in aqueous solution in the form of dichromate (Cr2O7−), hydrochlorate (HCrO4−), or chromate (CrO42−).22 These anionic species are usually not adequately adsorbed by negatively charged soil particles due to the nonrepulsive electrostatic interaction. Therefore, Cr(VI) is capable of mobility and prevails only in aqueous solutions. Most chromite mine discharge water contains higher Cr(VI) concentration and electroplating effluent and ferrochrome, and leather tanning industries contain even higher Cr(VI) concentrations than the permissible limit.23 Chromium compounds, of some significance considering human intake of the metal, can be used as an additive in water to prevent corrosion in industrial and various cooling systems. Opinions suggest that this utilization could cause a significant amount of environmental contamination from industrial emissions. As we will elaborate later, chromium salt utilization as a passivation agent on tin plate sometimes causes high levels of contamination in tinned foodstuffs. Although water may contain chromium, especially if it is affected by industrial emissions, it is not likely to account for impactful contributions to dietary intake of the element. Chromium's common form in drinking water is Cr(VI) and it is more soluble than hydrated Cr(III) oxide.22
Conditions have been documented in which residents in the vicinity of a chromium-contaminated site were potentially exposed to Cr(VI) in their drinking water at levels up to 10 ppm.24 Whether prolonged exposure to low concentrations of Cr(VI) in drinking water can lead to cancer is directly related to the current concern about the dangers of environmental exposure to Cr(VI). Subsequent poor disposal practices, such as dumping of Cr(VI) into unlined ponds, have led to Cr seepage into waterways and consequent contamination of irrigation and drinking water.25 Chromium compounds’ emissions in water and air are mainly caused by chemical manufacturing industries (rubber, dyes, pharmaceuticals, and plastic products). In humans, sufficient evidence exists for the carcinogenicity of Cr(VI) compounds which are found in chromate production, chromate pigment production, and chromium plating industries.26 Extremely toxic contaminants like As(V), Cr(VI), and Pb(II) were found in the chrome-tanning process of animal skin to produce leather; these industries were located in Riyadh and Jeddah, in Saudi Arabia, industrial cities.27 Chromium enters the oceans in two-way riverine discharge and atmospheric deposition. Almost no rivers flow into the Red Sea; hence, the major likely source of higher concentrations of Cr near Al Wahj and Jazan could be atmospheric deposition, activities of the Al Wajh Port, or wastewater streams from the local community.28 Chromium abundantly exists in the Earth's crust; its level in groundwater depends on the nature of the crust and the rocks in any given area. Accordingly, relatively high concentrations were detected in Makkah wells, though still below international and local standards.29 Different studies illustrated that Cr(VI) compounds can increase lung cancer risk, and that ingesting large amounts can cause ulcers, upset stomach, convulsions, and liver and kidney damage, and could even cause death.
Because of the undesirable inhibitory properties in terms of water pollution caused by dyes and heavy metals, elimination of these polluters from wastewater is key. Several methods including physical, mechanical, chemical, thermal, and biological methodologies are applied to remove these polluters from water systems. Ion exchange,34 precipitation,35 flocculation/coagulation,36 photo oxidation,37 phytoextraction,38 electrocoagulation,39 electrodialysis,40 irradiation,41 membrane separation,42 ultrafiltration,43 forward osmosis,44 and microorganisms and plants45–47 are applied to remove dyes and heavy metals from water bodies. Generally, wastewater treatment scenarios can be categorized into mechanical, chemical, physical, thermal, and biological methods as displayed in Fig. 1.
• Extract a large number of pollutants over a wide pH range
• Fast kinetics
• Easy handling and low space
• Regenerability
• Low cost
• No chemical sludge
Adsorption can be described as a metabolically independent and passive process that covers all interactions between any sorbate and biological matrix (biosorbent). It is crucial to many naturally occurring processes in different scientific disciplines, e.g., biotechnology, life sciences, and medical approaches.
To explore the main mechanism of adsorption, many strategies of characterization and theoretical calculation related to various models’ assumptions in terms of kinetics, thermodynamics, and isotherms were carried out. The adsorption interaction between water contaminants and different biosorbents takes place in two different scenarios: surface and interstitial sorption. In surface sorption, sorbate molecules displace from the aqueous solution to the biosorbent surface. Once polluters (molecules and/or ions) pass the biosorbent's surrounding boundary layer, they attach to the active sites present on its surface, displacing themselves from aqueous solutions. Dipole interactions, hydrogen bonding, or van der Waals forces are usually the cause of this type of sorption.55 On the other hand, the interstitial sorption mechanism differs, where polluters (molecules and/or ions) diffuse towards the biosorbents, entering the biosorbent pores (macro, meso, and/or micropores), and becoming biosorbed to the interior surface of the biosorbent.56
Using biosorbents to tackle different polluters from water and wastewater can be explained via electrostatic interactions. They have been established as major contributors to water contaminant adsorption. The presence of a plethora of functional groups (active sites) on the surface of the biosorbent elects them for efficiently capturing polluters from various water systems. The capacity of these functional groups is significantly affected by the medium's pH. The biosorbent surface charge and polluter's speciation are heavily reliant on the environmental pH. Lower pH values lead to protonation of various functional groups, which results in maintaining a positive charge on the surface of the biosorbent. As a consequence, electrostatic repulsion takes place, diminishing and/or inhibiting positively charged polluter sorption, i.e., metal cations. Conversely, electrostatic repulsion diminishes with increasing the pH of the medium, causing an increase in their adsorption in terms of electrostatic attraction phenomena. For example, negative-function moieties such as carboxylate (–COO−) and hydroxyl (–OH−) on the biosorbent surface maintain its negative charge, hence facilitating the binding of positive-charge polluters (molecules and/or ions). Contrarily, the holding of positive functional groups such as amine sites (–NH2) is considered as the main cause of the sorption of negative-charge water polluters.
The dissociation constant values (pKa) of biosorbent functional groups and their pHPZC values are relative to the solution's pH, hence influencing the capacity of biosorbent sorption. The pKa values of various functional groups such as carboxylic and phenol groups range from 3.5 to 5.5. This implies that the majority of these groups deprotonate in this pH working range and that avails enhancement of negatively charged sorption sites for the adsorption process.57 Additionally, the biosorbent (pHPZC) value greatly affects the adsorption process. The charge of the surface of the biosorbent is positive (pH of solution < pHPZC), while its surface charge is negative (pH of solution > pHPZC). Therefore, the elimination of polluters (molecules and/or ions), particularly those possessing positive charges, i.e., metal cations, is decreased when their pH is lower than pHPZC and increases when their pH is greater than pHPZC. This suggests the significant role of electrostatic forces during the process of adsorption.58
The adsorption process can be fathomed based on the ion exchange mechanism between the biosorbent and the studied polluters. Its contribution is illustrated through the replacement of protons from the exchangeable sites on the biosorbent surface with polluters, i.e., metal ions. Hydroxyl, carboxyl, and phenol groups facilitate the mechanism's operation, which can be attached to by polluter ions, i.e., divalent heavy metal ions, via two pairs of electrons and subsequently the release of two H+ and/or Na+ into the solution.59 The pH of the solution affects the ion exchange mechanism: in an acidic medium, the increase in the H+ ions leads to a competition with positively charged polluters (molecules and/or ions) leading to them being sorbed onto the biosorbent. Meanwhile in a basic medium, OH− ions’ increase compels negatively charged polluters (molecules and/or ions) to be sorbed on the sorption sites. For example, the possible ion exchange mechanism between polluters such as metal cations and exchangeable protons from the adsorbent surface can be illustrated as shown in the following equations:
Adsorbent⋯H+ + (metal ion)+ → Adsorbent⋯metal ion+ + H+ |
Adsorbent⋯Na+ + (metal ion)+ → Adsorbent⋯metal ion+ + Na+ |
Adsorbent⋯Ca2+ + (metal ion)+ → Adsorbent⋯metal ion+ + Ca2+. |
Formation of surface complexes (complexation) involves the interaction of polluters, i.e., metal ions, with oxygen donor atoms from oxygen-containing functional groups (coordination) associated with the release of protons and formation of surface complexes. The ligand tendency for forming metal complexes significantly depends on the metal classification according to their chemical characters including the Hard–Soft-Acid–Base principle (HSAB). Metal ions (cations) can bind with the biosorbent surface through inner- or outer-sphere complex mechanisms by the covalent bond chemically established between the metal and the oxygen atom (electron donor) or by cations nearing the negative groups present on the surface (critical distance) linked to the presence of at least one water molecule between the cation and base, respectively.53
Isotherm | Nonlinear form | Linear form | Theoretical assumption | Ref. |
---|---|---|---|---|
Langmuir | Assumes monolayer adsorption on a homogeneous surface with a definite number of adsorption sites | 60 | ||
Freundlich | q e = KFC1/ne | Suitable for a heterogeneous surface and can be applied to multilayer adsorption | 61 | |
Dubinin–Radushkevich | q e = QDRe−KDRε2 | lnqq = lnQDR − KDRε2 | Mainly used to determine the adsorption mechanism (physical or chemical adsorption) with the mean free energy on a heterogeneous surface | 62 |
Temkin | Assumes that there is a linear decrease in the heat of adsorption | 63 | ||
Flory–Huggins | Indicates the spontaneous nature and feasibility of the adsorption process | 64 and 65 | ||
Hill | Follows the hypothesis that ligand binding sites of a macromolecule can affect other binding sites of the same macromolecule | 66 | ||
Redlich Peterson | A combination of the Langmuir and Freundlich models. It can be used over a vast range of concentrations and can be used in both homogeneous and heterogeneous systems | 67 | ||
Sips | Used for predicting heterogeneous adsorption systems in a wide range of adsorbate concentration. It approaches the Freundlich isotherm at low concentration and the Langmuir isotherm at high concentration | 68 | ||
Toth | Widely used for inhomogeneous solid surfaces. It incorporates the effect of the interaction between the adsorbed substances as well | 69 | ||
Koble–Corrigan | Incorporated both the Langmuir and Freundlich isotherm models for representing the equilibrium adsorption data. It has an exponential dependence on concentration in the numerator and denominator. It is usually used with heterogeneous adsorption surfaces | 70 | ||
Khan | — | A general model for adsorption of a biadsorbate from pure dilute equation solutions. | 71 | |
Radke–Prausnitz | — | More preferred in most adsorption systems at low adsorbate concentration. At low adsorbate concentration, this isotherm model reduces to a linear isotherm, while at high adsorbate concentration it becomes the Freundlich isotherm | 72 | |
Frenkel–Halsey-hill | — | Mainly depending on the fractal dimension of the surface and on whether the dominant force is the substrate potential (van der Waals wetting, low coverage) or the film-vapor surface tension (capillary wetting, high coverage) | 73–76 | |
MacMillan Teller | — | Derived from the BET model including the effects of surface tension in layers formed in the adsorbent | 77 |
Kinetic model | Nonlinear form | Linear form | Theoretical assumption | Ref. |
---|---|---|---|---|
Pseudo-first order | q t = qe[1 − e−k1t] | Based on the sorption capacity and removal rate change with time being directly proportional to the difference in saturation concentrations (physical sorption) | 78 | |
Pseudo-second order | Investigates kinetic behavior in the case of a chemical reaction being the rate controlling step (chemical sorption) | 79 | ||
Intraparticle diffusion | — | q t = kit0.5 + X | Determines the rate controlling step of the sorption process (film diffusion and/or intraparticle diffusion) | 80 |
Elovich equation | Based on the difference in the sorption energies of sorbent active sites which resulted from the heterogeneous nature of the binding sites | 81 |
To acquire a high-efficiency and low-cost adsorbent for arsenic removal from water, a nanostructured Fe(III)–Cu(II) binary oxide was devised via the facile coprecipitation method.83 Various techniques including BET surface area measurement, powder XRD, SEM, and XPS were used to characterize the synthetic Fe(III)–Cu(II) binary oxide. Adsorption experiments took place to assess the adsorption kinetics, isotherms, and pH edge and regeneration of spent adsorbent. The results indicated that the Fe(III)–Cu(II) binary oxide with a Cu:Fe molar ratio of 1:2 performed excellently in displacing As(V) and As(III) from water, and the maximal adsorption capacities for As(V) and As(III) were 82.7 and 122.3 mg g−1 at pH 7.0, respectively.
It was attempted to synthesize cupric oxide nanoparticles (CuONPs) in a green effective way through utilizing lemon juice extract as a bioreductant.84 The CuONPs were applied for Cr(VI) displacement from water through the method of adsorption. The experimental parameters were optimized through the Box–Behnken Design (BBD) of response surface methodology for the optimum response. The adsorption equilibrium data were well fitted using the Freundlich isotherm model and the kinetics was explained through a pseudo-second-order kinetic model. The entire process proved to be cost effective, spontaneous, and exothermic. The linear approach for analyzing the isotherm as well as kinetic parameters was found to be more appropriate than the nonlinear approach. The experimental results showed that CuO nanoparticles can be good alternatives for the removal of Cr(VI) from aqueous solutions.
Another type of CuONPs has been effectively created via a green-effective route through employing lemon juice as a bioreductant.85 The green method for the synthesis of nanoparticles was cost effective, safe, and eco-friendly. The synthesized CuONPs were characterized by different analytical techniques, e.g., UV-visible, XRD, FTIR, SEM, and TEM analyses. This study revealed that the synthesized CuONPs were excellent solutions for detoxifying water of Cr(VI) through adsorption. The process of adsorption was found to be spontaneous and endothermic; the equilibrium during adsorption was well explained through pseudo-second-order kinetics.
Synthesizing copper oxide nanoparticles through the reactive magnetron sputtering technique and their adsorptive removal properties for Pb(II) ions from aqueous solution were investigated.86 Various adsorption parameters, e.g., the solution pH, adsorbent dose, initial metal ion concentration, equilibrium contact time and temperature were evaluated for removing maximum Pb(II) ions. The optimal prerequisites were found to be, respectively, 6, three hours, and 2 g L−1 for a 50 mg L−1 Pb(II) ion concentration. The adsorption kinetics followed the pseudo-second-order kinetic model, indicating that the adsorption was controlled through the chemisorption process. The adsorption isotherm follows a Langmuir isotherm with a maximum adsorption capacity of 37.03 mg g−1. The ΔS and ΔH values were found to be positive, indicating the endothermic nature of the adsorption process, while the negative value of the Gibbs free energy (ΔG) denotes that Pb(II) adsorption was spontaneous. The study revealed that the synthesized CuO nanoparticles may be a lucrative solution for removing Pb(II) ions from aqueous solutions.
A simple one-step approach for devising copper ferrocyanide-embedded magnetic hydrogel beads (CuFC-MHBs) was designed,87 and the beads were applied to effectively remove Cs and then magnetically separate it from water. The resulting CuFC-MHBs illustrated the effectiveness of Cs removal performance with a high Kd value of 66780 mL g−1 and exemplary structural stability without the release of CuFC for at least one month; it was effectively separated from water by an external magnet. Additionally, the CuFC-MHBs selectively adsorbed Cs with high Kd values in the presence of different competing ions, such as in simulated groundwater (24500 mL g−1) and seawater (8290 mL g−1). Also, they maintained their Cs absorption capacity in a wide pH spectrum, from 3 to 11. The convenient fabrication method and effective removal of Cs from various aqueous media showed that CuFC-MHBs show massive potential for practical application in the decontamination of Cs-contaminated water sources resulting from radioactive liquid waste and nuclear accidents in different nuclear industry fields.
The potential uses of ZnO and its microparticles and TiO2 nanoparticles to rid groundwater of arsenic were evaluated.89 The outcomes showed that the arsenic adsorption efficiency surged in proportion with the reduction of the adsorbents’ particle size. Upon using ZnO and TiO2 nanoparticles, their adsorption capacities were 0.85 and 0.99 mg g−1, respectively. TiO2 nanoparticles showed better adsorption ability for arsenic than that of ZnO since the first had a smaller average particle size and larger surface area. In aqueous solution, metal oxide nanoparticles have a hydroxyl surface that allowed the physisorption of incoming HAsO42− ions via hydrogen bonding; this resulted in a better arsenic adsorptive capacity.
Meng et al. (2018)90 developed a one-pot method to remove CrO42− from wastewater by forming Zn Al-Layered Double Hydroxides (ZnAl-LDH). By this process, the efficiency of Cr(VI) removal was nearly 100% and reached the theoretical maximum of the ZnAl-LDH removal capacity. In addition, mixed contaminants of Cr(VI) and dyes in wastewater can be separated from each other by controlling the addition of Zn/Al solution using this one-pot method.
A ZnO-tetrapod-activated carbon (ZnO-T/AC) nanocomposite was synthesized as an adsorbent for efficient decontamination of Cr(VI) from an aqueous medium.91 The reduction of Cr(VI) using the ZnO–AC adsorbent was influenced by the pH value, contact time and dose of the adsorbent. The maximum reduction efficiency of Cr(VI) to Cr(III) by AC, ZnO, and ZnO–AC adsorbents occurred at pH 2.0. At optimum pH 2, AC, ZnO-T, and ZnO-T/AC showed 43%, 55%, and 97% Cr(VI) reduction, respectively. The pseudo-second-order model and intraparticle diffusion and Elovich models were used to determine the adsorption capacity and the kinetic rate constant, and fitted well for all the adsorbents.
Sulfidated zerovalent iron (S-ZVI) is garnering more attention owing to its easy-to-prepare and highly-reactive nature with contaminants.93 The results showed that sulfidation of zerovalent iron (ZVI) could significantly increase the Cr(VI) and Sb(III) sequestration simultaneously. During the sequestration of these pollutants, Cr(VI) was diminished into Cr(III) and Sb(III) was oxidized into Sb(V); both processes occurred simultaneously in this system. The sequestration rate of Cr(VI) and Sb(III) increased with the pH surging from 4.0 to 6.0 but diminished when the pH continued to rise to 8.0. More surface Fe(II) regenerated at pH 6.0, which significantly contributes to Cr(VI) sequestration.
Mubarak et al. (2018)94 prepared Mg/Fe-layered double hydroxide (MF–LDH) hollow nanospheres by a simple thermal method. After calcination at 400 °C, the MF-LDH hollow nanospheres were converted into the corresponding oxide. The nanospheres showed remarkable efficiency in removing As(V) and Cr(VI) ions, with maximum adsorption capacities of 178.6 mg g−1 for As(VI) and 148.7 mg g−1 for Cr(VI). Complete heavy metal removal (∼99.9%) from wastewater up to the drinking water level (WHO standards) was achieved in 20 min and 10 min for As(VI) and Cr(VI), respectively. This material should facilitate practical applications in cost-effective wastewater purification.
Siderite (FeCO3), an iron(II)-containing mineral, exists in many anaerobic sediments and groundwater systems. Siderite is an effective material for removing Cr(VI) in aqueous conditions with a maximum sorption of 81 mg g−1 at pH 5.95 The outcomes denote that within anoxic aqueous parameters, Cr(VI) can be detoxified by siderite provided that these coupled sorptions and redox reactions are being controlled by the initial Cr(VI) concentrations and pH.
Graphene-coated iron oxide nanoparticles (GCIO) were devised via the microemulsion method for removing Cr(VI) ions from aqueous solutions.96 A swift and remarkable Cr(III) and Cr(VI) ion removal efficiency is exhibited by aqueous solution-prepared GCIO, which is 99.13% with 0.06 g GCIO in 100 ppm adsorbate solution as confirmed by EDX analysis. The adsorption process follows pseudo-second-order kinetics with an R2 value of 0.97 and the Freundlich model isotherm with a high adsorption capacity. The Cr(VI) removal surges with the increase in temperature, revealing the endothermic nature of the process of adsorption. GCIO also shows a high intraparticle diffusion value with an R2 value of 0.98. GCIO maintains remarkable stability in an extremely acidic environment when treated with concentrated acid for 24 hours as well as a reusable nature to the same extent for adsorption of Cr(VI).
A multifunctional adsorbent was developed by preparing a composite of Fe2O3 and ZnO.97 The prepared composite removed Cr(VI) efficiently. The data of adsorption corroborated well with the Freundlich isotherm model and pseudo-second-order kinetic model with R2 values close to 1. The 1:1 ratio of Fe2O3 and ZnO provided the optimum sorption properties to the sorbent. At this ratio, the maximum adsorption capacity of the adsorbent was 36.50 mg g−1.
Mechano-chemically sulfidated microscale zerovalent iron (S-mZVIbm) is a potential remedy for groundwater for the removal of Cr(VI).98 Removing Cr(VI) by S-mZVIbm is mainly a process of a chemisorption nature and its kinetics was well fitted by a pseudo-second-order model. Alkaline pH halted the removal of Cr(VI) while dissolved oxygen slightly depressed the removal of Cr(VI). The high extent of removal of Cr(VI) by S-mZVIbm compared to its unsulfidated contemporary mZVIbm is a result of its enhanced surface area.
Another nanoscale zerovalent iron (nZVI) sorbent with a considerably high surface area (182.97 m2 g−1) with chain-structure morphology was synthesized and well characterized.99 The as-prepared nZVI can entirely displace Cr(VI) under anoxic conditions after a reaction of 20 min; on the other hand, only 43% of Cr(VI) was removed after a reaction of 60 min under oxic conditions. Notably, nZVI demonstrates a remarkable removal capacity of Cr(VI) (123.85 mg g−1) as well as a quick removal rate (0.017 g mg−1 min−1). Moreover, the as-prepared nZVI exhibited swift displacement of traced Cr(VI) from Cr-spiked drinking water or actual Cr-contaminated lake water. Cr(VI) removal experimental results from simulated Cr-contaminated tap, drinking, and actual lake water show the effectiveness of the as-prepared nZVI as a potential material for the quick removal of traced Cr(VI) from Cr-spiked drinking water or actual Cr-contaminated lake water.
After being analyzed and characterized, nanoadsorbents of core–shell bimagnetic nanoparticles (CoFe2O4@γ-Fe2O3) with two different mean sizes were applied as potential sorbents for removing Cr(VI) from aqueous solutions via magnetic-assisted chemical separation.100 The adsorption data checked out with the Freundlich model, denoting that the Cr(VI) species create multilayers surrounding the nanoadsorbent surface. The utmost removal percentage was gauged at pH = 2.5 with an orbital shaking rate of 400 rpm, and the maximum adsorption capacities obtained were higher or close to those of other magnetic adsorbents dedicated to removing Cr(VI). The improved adsorption capacity of the sample according to smaller nanoparticles was related to its higher BET surfaces. The adsorption rapidly achieved equilibrium and the kinetic data adhered to the pseudo-second-order model, implying that the rate-controlling step of the process revolved around ionic interactions between the surface sites and the Cr(VI) species. The thermodynamic analysis yielded that the adsorption is endothermic and spontaneous. The process was found to be enhanced at higher temperatures and presented increased randomness at the surface/solution interface. The magnetic nanoadsorbents also demonstrated a high selectivity towards Cr(VI) and good renewability for reuse, establishing them as a greener material. The adsorption capability remained high in simulated wastewater. This process can be efficiently applied to other related heavy metals such as molybdenum and arsenic.
Jain et al. (2018)101 fabricated magnetite (Fe3O4) and magnetite/activated-carbon (Fe3O4/AC) through the coprecipitation method for removing Cr(VI), Cu(II), and Cd(II) ions from aqueous solution in batch mode. The optimal conditions for the removal of ions were as follows: pH = 2 for Cr(VI) desorption studies with 0.1 M HCl. It is stated that these nanoparticles can be regenerated effectively, allowing them to be reused for up to four adsorption–desorption cycles without any loss in mass.
A highly selective Fe3O4@Arg-PPy adsorbent was effectively synthesized by a one-pot chemical polymerization of pyrrole with arginine and was made commercially available as Fe3O4 NPs. The FTIR, VSM, XRD, and TEM analyses of the Fe3O4@Arg-PPy nanocomposite imply the successful implementation of arginine and Fe3O4 into PPy.102 The batch experimental results showed a highly pH-dependent removal of Cr(VI) at the optimum pH value of 2. The process of adsorption was Langmuir modeled with adsorption capacities of 294.11, 322.58, 370.37, and 416.67 mg g−1 at 15, 25, 35, and 45 °C, respectively. The Cr(VI) removal kinetics via Fe3O4@Arg-PPy conformed to the pseudo-second-order model, and the determined rate constant was 6.67 × 10−3 g m−1 min−1 at room temperature for a solution of an initial Cr(VI) concentration of 50 mg L−1 and pH 2. The thermodynamic parameters that were calculated were as follows: ΔG° −4.76 kJ mol−1 at 25 °C, ΔH° +76.16 kJ mol−1, ΔS° + 0.27 kJ mol−1 K−1 and Ea 49.91 kJ mol−1, showing an endothermic, spontaneous, and chemisorption process.
A Microsized Granular Ferric Hydroxide (mGFH) adsorbent for chromate removal in competition with ions present in drinking water was elaborated.103 The results show a high dependency on the pH value with increasing adsorption in correspondence to decreasing pH values. Competing ions present in drinking water were tested for interfering effects on chromate adsorption. Bicarbonate was identified as the main inhibitor of chromate adsorption. Sulfate, silicate, and phosphate also decreased chromate loadings, while calcium enhanced chromate adsorption. The adsorption equilibrium was reached after 60 min for particles smaller than 63 mm, while 240 min was required for particles from 125 mm to 300 mm. The adsorption kinetics in single-solute systems could be modeled using the Homogeneous Surface Diffusion Model (HSDM) with a surface diffusion coefficient of 4 × 10−14 m2 s−1.
Wen et al. (2019)104 successfully synthesized nanosized ordered Magnetic Mesoporous Fe–Ce bimetal oxides (nanosized-MMIC) with a highly well-ordered interconnected mesostructure. This nanosized-MMIC exhibited remarkable adsorption capabilities for As(V), Cr(VI), and acid orange 7 (AO7) and the corresponding calculated maximum adsorption capacities of the materials were 111.17, 125.28, and 156.52 mg g−1, respectively. The removal of As(V) and Cr(VI) by nanosized-MMIC was slightly dependent on the ionic strength, yet it was highly solution-pH dependent; the coexistent silicate and phosphate ions competed excellently with As(V) and Cr(VI) for the adsorption active sites. The mechanisms implied that As(V) and Cr(VI) formed inner-sphere complexes on the nanosized-MMIC interface through surface complexation and electrostatic interaction. Nanosized-MMIC is anticipated to be a potential excellent nano-adsorbent with high potential for application for removing coexisting organic dyes and toxic heavy metals in specific wastewater treatment. The adsorption capacity of some selected iron incorporating adsorbents are reported in Table 4.
No. | Magnetic adsorbents | Adsorbate | Adsorption capacity (mg g−1) or removal efficiency (%) | Ref. |
---|---|---|---|---|
1 | Fe3O4@thiourea-formadehyde polymer | Cr(VI), As(V) | Cr(VI) (4.28 mmol g−1), As(V) (1.97 mmol g−1) | 105 |
2 | MgFe2O4 | Co(II), Mn(II), Ni(II), Cu(II) | Co(II) (2.30 mmol g−1), Mn(II) (1.56 mmol g−1), Ni(II) (0.89 mmol g−1), Cu(II) (0.46 mmol g−1) | 106 |
3 | Dried ZVI NPs | Pb(II) | 807.23 mg g−1 at pH 6 | 107 |
4 | ZnFe2O4@NH2–SiO2@polydiphenylmethane diisocyanate@dithizone | Pb(II) | 267.2mg g−1 | 108 |
5 | ZVI/activated carbon nanotubes | Te(II) | 800 mg g−1 at pH 4.7 | 109 |
6 | Ammonium-pillared montmorillonite/CoFe2O4/calcium alginate | Cs+ | 86.46 mg g−1 | 110 |
7 | ZVI-Coffee grounds | Pb(II), Cd(II), As(III), As(V) | Pb(II) 164.1 mg g−1 (1 h), Cd(II) 112.5 mg g−1 (24 h), As(V) 9.3 mg g−1 (1 h) As(III) 23.5 mg g−1 (1 h) | 111 |
8 | Clinoptilolite/CoFe2O4 | Sr(II) | 20.58 mg g−1 | 112 |
9 | ZVI/chitosan | U(VI) | 591.72 mg g−1 at pH 6 | 113 |
10 | MnFe2O4 | U(VI), Eu(III) | U(VI) 119.90 mg g−1, pH 5 Eu(III) 473.93 mg g−1, pH 7 | 114 |
11 | ZVI NPs | Cu(II) | 343 mg g−1 | 115 |
12 | Hydroxyapatite/NiFe2O4 | 152+154Eu 160Tb | Eu(III) (137.35 mg g−1) Tb(III) (130.43 mg g−1) | 116 |
13 | ZVI/alginate | Pb(II) | 581.7 mg g−1 (15 min) | 117 |
14 | Ni0.6Fe2.4O4 | U(VI) | 189.04 mg g−1 (2 h) | 118 |
15 | ZVI/biochar | Cr(VI) | 35.30 mg g−1 | 119 |
16 | NiFe2O4/MnO2 | Pb(II) | 85.78 mg g−1 | 120 |
17 | ZVI/polyaniline/attapulgite | Cr(VI) | 86.56 mg g−1 | 121 |
18 | MgFe2O4/biochar | PO43− | 163.02 mg g−1, pH 3 | 122 |
19 | ZVI/ZVAl | Cr(VI), Cd(II), Ni(II), Cu(II), Zn(II) | 99.5% (300 h) | 123 |
20 | Poly(m-phenylenediamine)/GO/NiFe2O4 | Cr(VI) | 502.5 mg g−1 at pH 3 | 124 |
21 | ZVI NPs | Cd(II), Cu(II), Ni(II), Pb(II) | Cd(II) (79.33–102.00) mg g−1 Cu(II) (111.11–142.85) mg g−1, Ni(II) (107.30–137.96) mg g−1, Pb(II) (110.97–142.68) mg g−1 | 125 |
22 | NiFe2O4-nitrogen-doped mesoporous carbon | Hg(II) | 476.2 mg g−1 | 126 |
23 | ZVI/fly ash | Pb(II), Cr(VI) | Pb(II) 78.13 mg g−1, Cr(VI) 15.70 mg g−1 | 127 |
24 | Ni0.5Zn0.5Fe2O4 | Ag(I) | 243.90 mg g−1 | 128 |
25 | Bentonite/ZVI | Pb(II), Cu(II), Zn(II), Ni(II) | Pb(II) (3.1 mg g−1), Cu(II) (0.6 mg g−1), Zn(II) (2.8 mg g−1), Ni(II) (2.4 mg g−1) | 129 |
26 | Bacterial cells/sawdust/MnFe2O4 | As(III) As(V) | As(III) (87.573 mg g−1), As(V) (88.990 mg g−1) | 130 |
27 | ZVI/magnetite carbon | U(VI) | 203.94 mg g−1 | 131 |
28 | GO/NiFe2O4 | Pb(II) Cr(III) | Pb(II) (25.0 mg g−1), Cr(III) (45.50 mg g−1) | 132 |
29 | ZVI/polyaniline-graphene aerogel | U(VI) | 350.47 mg g−1 at pH 5.5 | 133 |
30 | CoFe2O4–G & NiFe2O4–G | Pb(II), Cd(II) | Pb(II) is 142.8 and 111.1 mg g−1 at pH 5 and 310 K for CoFe2O4–G & NiFe2O4–G; while for Cd(II) it was 105.26 and 74.62 mg g−1 at pH 7 and 310 K | 134 |
31 | ZVI/activated carbon | As(V) | 100% after 2.5 h | 135 |
32 | α-Fe2O3 | Hg(II) | 1.16 mg g−1 | 136 |
33 | Fe3O4/bone char/chitosan | As(V) | 112 μg g−1 | 137 |
34 | γ-Fe2O3 | Cd(II), Ni(II), Co(II) | Cd(II) (94.33 mg g−1), Ni(II) (86.206 mg g−1), Co(II) (60.60 mg g−1) | 138 |
35 | Fe2O3–Al2O3 | Cu(II), Pb(II), Ni(II), Hg(II) | Cu(II) (4.98 mg g−1), Pb(II) (23.75 mg g−1), Ni(II) (32.36 mg g−1), Hg(II) (63.69 mg g−1) | 139 |
36 | γ-Fe2O3 | Pb(II), Zn(II), Cd(II) | Pb(II) (10.55 mg g−1), Zn(II) (4.79 mg g−1), Cd(II) (1.75 mg g−1) | 140 |
37 | ZVI/zeolite | Cd(II), Pb(II) | Cd2+ (63.14 mg g−1), Pb2+ (154.61 mg g−1) | 141 |
38 | Fe3O4/PANI | Cr(VI) | 200 mg g−1 | 142 |
39 | Fe2O3@GO | Pb(II) | 303.0 mg g−1 | 143 |
40 | γ-Fe2O3/TiO2/PVA/alginate beads | Ba(II) | 99% in 150 min at pH 8 | 144 |
41 | γ-Fe2O3@chitosan | Cd(II) | 15.2 mg g−1 | 145 |
42 | Chitosan/Cu(OH)2, chitosan/CuO | Cr(VI) | 1.4 mmol g−1, 2.6 mmol g−1 | 146 |
43 | Polyacrylonitrile/α-Fe2O3 | As(V) | 82.2 mg g−1 at 25 °C | 147 |
44 | Fe3O4/MgAl-layered double hydroxide | Co(II) | 95.8% at pH 8 | 148 |
45 | Chitosan/PVA/ZVI | As(V) | 200 mg g−1 | 149 |
46 | ZVI/diethylenetriamine/2-pyridinecarboxaldehyde | Co(II), Cu(II), Zn(II), Cd(II), Hg(II), Pb(II) | Co(II) 2600 μmol g−1, Cu(II) 4750 μmol g−1, Zn(II) 5600 μmol g−1, Cd(II) 4000 μmol g−1, Hg(II) 5200 μmol g−1, Pb(II) 5050 μmol g−1 | 150 |
47 | ZVI/MnO2 | As(V) | 16.4 mg g−1 | 151 |
48 | Fe3O4/alginate beads | La(III) | 1.8 mmol g−1 | 152 |
49 | γ-AlOOH/α-Fe2O3 | Cr(VI) | 4.17 mg g−1 | 153 |
50 | Chitosan/clay/Fe3O4 | Cu(II), As(V) | Cu(II) (17.2 mg g−1), As(V) (5.9 mg g−1) | 154 |
51 | Fe3O4/poly-p-phenylenediamine-thiourea-formaldehyde | As(V) | 99.04 mg g−1 | 12 |
52 | Fe3O4/GO/beads | Cr(VI), As(V) | Cr(VI) (80%), As(V) (99%) | 155 |
53 | Fe3O4@cyclodextrin | Eu(III) | 95% at pH 8 | 156 |
54 | Fe3O4/chitosan | Cr(VI) | 231.77 mg g−1 | 13 |
55 | Functionalized nanosilica | Cu(II) | 386.4 mg g−1 | 157 |
56 | Diethylenetriamine/GO/Fe3O4 | Cr(VI) | 123.4 mg g−1 (40 s) | 158 |
57 | Chitosan-based magnetic composite | Diclofenac sodium tetracycline hydrochloride | 164 mg g−1, 40.2 mg g−1 | 159 |
58 | Zeolitic imidazolate frameworks confined in polystyrene | Benzotriazole | 210.4 mg g−1 | 160 |
59 | Poly(N-isopropyl acrylamide) grafted chitosan/Fe3O4 composite particles, [CN-MCP] | Nonylphenol | 166.7 mg g−1 | 161 |
60 | Aromatic ring-functionalized chitosan magnetic composite | Norfloxacin, tylosin, diclofenac sodium | 327.5 mg g−1, 843.5 mg g−1, 438.6 mg g−1 | 162 |
Mikhaylov et al. (2018)164 synthesized Al/Fe oxide–oxyhydroxide composite powders through a hydrothermal method applying metal sulfate solutions and urea as precursors. The maximum specific surface area (190 m2 g−1) is reached for the synthesized samples synthesized from solutions with [Al3+]:[Fe3+] = 1:0 and 1:1. However, sulfate ions present on the surface of Al-containing samples inhibit the adsorption of Cr(VI). The maximum capacity for adsorption in the batch regime (3.66 mg g−1) was for a sample synthesized from solutions with [Al3+]:[Fe3+] = 1:6, which had a high (though not maximal) surface area (102 m2 g−1), a low-sulfate content, and maximum zeta potential. Sorption experiments illustrated that the mean free adsorption energy was lower than 8 kJ mol−1 for the entirety of the samples; this indicates that the process of adsorption was physical and the possibility of Cr(VI) desorption and adsorbent regeneration.
Another type of metallic adsorbent based on Apatitic Tricalcium Phosphate powder (ATrPh-105) was devised and analyzed by different techniques of characterization.165 At 60 °C and under acidic conditions, a higher adsorption capacity (527.19 mg g−1) towards Cr(VI) was recorded. The kinetic data of the adsorption of Cr(VI) onto ATrPh-105 adhered to pseudo-second-order kinetics. The results of isotherms and thermodynamics studies affirmed that Cr(VI) adsorption onto ATrPh-105 is endothermic, which happens to fit the Langmuir isotherm as well. According to the thermodynamic results of the sorption of Cr(VI) onto ATrPh-105, the process was controlled by a chemical reaction. Additionally, ATrPh-105 illustrated Cr(VI) high sorption efficiency (90.22%) following three cycles of adsorption–regeneration. This finding confirms the potential of ATrPh-105 as an efficient adsorbent with room for utilization in various environmental applications.
Dutta and Nath (2018)167 synthesized a SiO2/carbon nanocomposite (SiO2/CCNC) and nanoporous carbon (CCNC) from a simple, low-cost, and industry-scalable method from corn cob biowaste. The higher BET specific surface area of SiO2/CCNC (715.22 m2 g−1) compared with CCNC (430.17 m2 g−1) allows for enhanced adsorption efficiency. The uptake of Cr(VI) ions was also tested in the presence of interfering salts such as NaCl/Na2SO4. The effects of temperature, pH, and concentration of the dye on the efficiency of sorption of the SiO2/CCNC nanocomposite and CCNC for MB dye were evaluated. The adsorption process kinetics adhered to the pseudo-first-order model for CCNC and a pseudo-second-order model for the SiO2/CCNC nanocomposite. The experimental data acquired for SiO2/CCNC and CCNC adhere to Bangham's pore diffusion model, indicating pore diffusion as the rate-controlling step during MB adsorption. The isotherm data match the Freundlich model for SiO2/CCNC and CCNC, which is a telltale of multilayer sorption. Hence, it can be deduced that SiO2/CCNC is a potential low-cost adsorbent compared with commercial activated carbon for removing pollutants like MB dye and U(VI)/Cr(VI) ions from water and wastewater.
Carboxyl-group-functionalized mesoporous silica (CFMS) prepared by a one-pot cocondensation method was employed for the Solid Phase Extraction (SPE) of chromium species.168 The captured Cr(III) can be eluted by using HNO3 and detected by ICP-MS, and the concentration of Cr(VI) was the difference between the total chromium and Cr(III). This protocol has been successfully applied to detect inorganic chromium species in environmental water (rain and lake and river water) with recoveries between 91.9% and 103%. Therefore, this developed method exhibited good potential for the speciation of chromium in environmental water samples.
Elwakeel et al. (2018)170 prepared a composite material (PAN-Na-Y-zeolite) through the polymerization of acrylonitrile in the presence of Na-Y-zeolite. The composite was made usable through amidoximation as a result of the reaction of hydroxylamine on nitrile groups of the composite. The adsorbent (APNa-Y-zeolite) was characterized by FTIR spectrometry, XRD diffraction, SEM, TGA, zetametry, and BET analysis. The adsorption properties of APNa-Y-zeolite were evaluated for the recovery of Cu(II), Cd(II), and Pb(II) from synthetic solutions prior to being tested for purity standards of drinking water. The properties of adsorption were profiled through the study of the effect of pH, adsorption isotherms, and uptake kinetics. The PSORE adhered to the kinetic profiles. Langmuir, Freundlich, and Sips equations were used to model adsorption. Variation of temperature was used as a benchmark to evaluate thermodynamic parameters. While the adsorption of Cu(II) and Cd(II) was endothermic, Pb(II) recovery was exothermic. Metal ions were effectively desorbed using 5 M HCl solutions. High concentrations of NaCl scarcely affect the sorption performance.
Spent C. monacanthastem (CS) is a copiously available agro-waste material that was experimented upon for adsorbing noxious pollutants from water systems. Toxic Cr(VI) adsorption has been verified by EDX and FTIR techniques. The adsorption behavior of the adsorbates was highly correlated with the pH of the solution.172 The maximum adsorption of Cr(VI) was recorded at 2.0, and the maximum monolayer adsorption capacity (Qm: mg g−1) of Cr(VI) was recorded to be 67.88 mg g−1. The thermodynamic evaluation confirmed the endothermic and spontaneous displacement of adsorbate ions from the aqueous phase. On the other hand, the desorption assessment showed that the adsorbate was efficiently removed from the adsorbent's surface with greater adsorption uptake (%) after five cycles. The use of CS–AC as an adsorbent for Cr(VI) was highly economic and competent, indicating potential use for detoxification of water.
The utilization of magnetic activated carbon modified with Poly Amidoxime (PAMC) as a potent methodology for chromium and thallium adsorption was studied.173 Activated carbon was devised from waste rubber tires, subsequently altered with magnetic properties, and then changed with poly amidoxime to provide more functional groups and thus improve the sorption efficiency. Tests were also carried out on real wastewater. With easy separation and high efficiency of removing chromium and thallium from water, besides reliable regeneration without significant loss in the capacity of adsorption, even after multiple cycles, the material exhibited encouraging potential as an adsorbent for treating water. The obtained result showed that the adsorbent works in a broad spectrum of pH from 3 to 11 and at an initial concentration of 1 to 20 ppm with an optimal shaking speed of 200 rpm. Adsorbents were easily displaced from the liquid after adsorption of the toxic pollutants from water; this indicates that PAMC has a promising method as an adsorbent for treating water.
Electrospinning allowed the facile synthesis of a mechanically stable nanofiber network, while hydrothermal treatment attained an α-Fe2O3 nanostructure surface coating that expanded the reactive surface area allocated for the uptake of dissolved metals.174 Composite nanofibers of polyacrylonitrile (PAN) with embedded hematite (α-Fe2O3) nanoparticles were synthesized via a single-pot electrospinning synthesis. A core–shell nanofiber composite was also prepared through the subsequent hydrothermal growth of α-Fe2O3 nanostructures on embedded hematite composites. These materials were tested for Cu(II), Pb(II), Cr(VI), and As(V) sorption. Although the uptake of Cu(II), Pb(II), Cr(VI), and As(V) was largely independent of the core–shell variables explored, metal uptake on embedded nanofibers was increased with α-Fe2O3 loading. Fig. 6 summarizes the incorporation of different metals into biosorbents.
Nanomineral-modified biochars show potential capacity for adsorbing and thus removing pollutants through combining the advantages of the porous structure of biochar and nanominerals’ unique properties.176 Therefore, nano ZnO/ZnS-modified biochar was created from the slow pyrolysis of zinc-contaminated corn stover resulting from the process of adsorption. Consequently, the nano ZnO/ZnS-modified biochar exhibited far better sorption capacity in terms of Pb(II), Cu(II), and Cr(VI) compared with the common biochar. The adsorption pattern complied with heterogeneous adsorbents, reflected by adhering to the Freundlich model. The thermodynamic outcomes implied that the adsorption process was spontaneous and endothermic. The method could be a dependable and highly lucrative approach to transforming heavy metal–polluted biomass to highly efficient nanomineral-modified biochar.
Biochar that was derived from Enteromorpha prolifera and magnetically modified was evaluated for its sorption properties towards Cr(VI).177 The pH and background ion intensity of the solution exhibited significant effects on Cr(VI) adsorption. Besides, the magnetic Enteromorpha prolifera derived biochar could be effectively recovered magnetically and regenerated with an alkaline solution. Therefore, this magnetic biochar derived from Enteromorpha prolifera has the potential to serve as a highly efficient adsorbent for water pollution control.
Biochar has redox activity that can be involved in environmentally relevant redox reactions. The electron transfer for the reduction transformation of Cr(VI) during its sorption by biochar was evaluated. Biochar extracted from peanut shells at 350 °C could impeccably displace Cr(VI) from solution, paired with reducing Cr(VI) to Cr(III), which was more apparent at strong acidity (pH 2).178 As a result of electron acceptance from lactate and/or biochar, Cr(VI) was reduced into CrOOH, which was verified by X-ray diffraction analysis. The outcome implies that biochar possesses potential for acting as an electron donor and shuttle for reducing Cr(VI) during the process of sorption, posing it as an alternative for detoxifying Cr(VI) from wastewater.
Conversion of carcinogenic Cr(VI) to less toxic Cr(III) has always been deemed an effective method to decontaminate Cr(VI)-contaminated water. However, the commonly utilized reducing agents such as nano zerovalent iron (nZVI) and its derivatives usually are unstable and pose the threat of yielding secondary contamination.179 To focus on these, nZVI was loaded on sludge-derived biochar to create a nZVI-BC composite for removing Cr(VI). Re-pyrolysis thus can be applied as an effective technology to diminish the environmental risks of exhausted adsorbent biochar for safe disposal. The adsorbent characterization showed that most Cr adsorbed on nZVI-BC was in the form of Cr(III) hydroxides or iron chromium hydroxides. The process of removal was extremely pH dependent. The adsorption kinetics was well modeled by the pseudo-second-order model and the intraparticle diffusion model. The Langmuir model illustrated the isotherms well and the maximum Cr(VI) removal capacity of nZVI-BC was 31.53 mg g−1 at pH 4 and 25 °C. The thermodynamic analysis yielded that the process of sorption was deemed spontaneous. Results from the risk analysis postulate that direct disposal of the postadsorption biochar threatens the environment. Re-pyrolysis of the exhausted adsorbent biochar, however, can optimally eliminate the bioavailable content of Cr and reduce its latent negative impact.
Biochar-supported nano-scale zerovalent iron (biochar-CMC-nZVI) stabilized by carboxymethyl cellulose (CMC) was devised and utilized for removing Cr(VI) from an aqueous solution.180 The outcome showed that 100 mg L−1 Cr(VI) could be displaced entirely by biochar-CMC-nZVI within 18 hours, at a dosage of 1.25 g L−1, and an initial pH value of 5.6.Cr(VI). Using biochar-CMC-nZVI to remove contamination is favored when lower pH is present. Sorption kinetics and isotherm results confirmed that the adsorption progress of Cr(VI) by biochar-CMC-nZVI is chemisorption of monolayers under optimal conditions. The pH value contributed immensely to Cr(VI) removal as the hydrogen ion could elevate the reduction of Cr(VI) to Cr(III) and thusly facilitate Cr(III) precipitation. These results showcase the biochar-CMC-nZVI composite potential as a green-effective, low-cost sorbent for removing Cr(VI) in the environment.
Kumarathilaka et al. (2016)181 evaluated the Cr(VI) reduction ability and removal using a stabilized nanozerovalent iron-graphene composite (NZVI-Gn). The composite was characterized by XRD, SEM, FTIR, BET, and pHpzc and NZVI was found to be homogeneously impregnated onto graphene sheets with a high surface area. The results revealed that the NZVI-Gn composite achieved excellent efficiency of removal within acidic conditions. It was suggested by the kinetic study that PSORE could illustrate the adsorption behavior of the NZVIGn composite so that the rate-limiting factor could be a chemisorption process. The Langmuir isotherm model provided a better correlation of the experimental equilibrium data, hinting that Cr(VI) adsorption onto a composite of NZVI-Gn could involve a process of a monolayer nature. Moreover, dual sorption and reduction followed by the immobilization process could elevate the removal of Cr(VI) immensely. Therefore, the preset NZVI-Gn composites are considered an admirable and efficient alternative and magnetically separable adsorbent for removing Cr(VI) from the environment.
A hierarchical porous carbon monolith with a 3D-interconnected structure was derived via resorcinol/melamine/formaldehyde as precursors.182 This hierarchical porous carbon maintains a high specific surface area of 1808 m2 g−1 and a well-developed porous structure, proved to be beneficial for removing Cr(VI). The experiments of adsorption show that the hierarchical porous carbon is an exemplary Cr(VI) adsorbent with an adsorption aptitude of 463 mg g−1 at pH 1. The developed porous carbon monolith maintains high selectivity for Cr(VI) ions in a coexisting system of Ni(II), Zn(II), and Cu(II). Therefore, this novel hierarchical porous carbon with controllable hierarchical porosity possesses good potential for water treatment application. Monolith strength improvement, adjustment of the carbon surface chemistry, and the performance in a more realistic water treatment system will be studied in our future work.
Hydroxyethyl cellulose and CuO nanoparticles making up a clay-alumina ceramic composite membrane (CuO/HEC) were fabricated for separation of Cr(VI) and Pb(II) from contaminated water.184 They were prepared by a slurry-casting method over a low-cost clay alumina ceramic substrate. It is believed that these CuO nanoparticle-containing clay-alumina ceramic composite membranes will be an effective solution for removing heavy metals from contaminated water. The prepared composite represents a reasonably high-performance novel ceramic UF membrane for the separation of toxic metal ions from mixed-salt solution. However, future study related to the use of the developed membrane for the treatment of real wastewater is necessary before implementation on a pilot scale.
Cellulose Aerogels (CA) and Zeolitic Imidazolate Framework-8 (ZIF-8) are combined into one highly functional aerogel for removing Cr(VI) from water through adsorptive means as a facile method of combining two emerging materials, by entrapping ZIF-8 particles into a flexible and porous CA (ZIF-8@CA).185 The adsorption capacity of the cellulose aerogel for Cr(VI) was significantly increased when ZIF-8 was added, and the Langmuir maximum adsorption capacity reaches up to 41.84 mg g−1. This kind of material may also be extended to be used as air filters and sensors, and for substrate-supported catalysis.
An efficient adsorbent of bacterial cellulose/poly(m-phenylenediamine) (BC/PmPD) was prepared for removing Cr(VI) through merging the merits of PmPD's high adsorption capacity and BC's ease of reclamation.186 The Cr(VI) BC/PmPD adsorption was ascribed to the Cr(VI) adsorption on protonated –NH– and –NH2 groups and the redox reaction of Cr(VI) to Cr(III) by the reduction of amine. Afterward, Cr(III) chelation on imino groups of PmPD took place. The calculated adsorption capacity of BC/PmPD for Cr(VI) was as high as 434.78 mg g−1. The Cr(VI) adsorption kinetics on BC/PmPD can be properly modeled by the pseudo-second-order kinetic model, placing the rate-controlling step as the chemical reaction. The findings in that study showed that BC/PmPD can be deemed as a potential adsorbent for removing Cr(VI) thanks to its efficient reclamation, high adsorption capacity, and good regeneration aptitude.
A composite adsorbent (PCA@AC) was successfully devised via cellulose (AC) as a substrate and poly(catechol tetraethylenepentamine-p-phenylenediamine) as a coating layer under a multitude of technical conditions.187 This composite maintains a remarkable performance for removing Cr(VI) from aqueous solutions, surpassing pure AC. The PCA@AC maximum adsorption capacity for Cr(VI) is obtained as 507.61 mg g−1 at 30 °C using the Langmuir model, surpassing pure AC. The removal ability of PCA@AC is attributed to the electrostatic interaction and reduction reaction of Cr(VI) to Cr(III) within the adsorption process. The composite adsorbent (PCA@AC) has great potential for Cr(VI)-wastewater purification with excellent chelating ability.
Singh et al. (2014)189 evaluated the effect of starch functionalization of iron oxide nanoparticles on their adsorption of Cr(VI) from aqueous solutions. Starch-Functionalized (SIONPs) as well as nonfunctionalized Iron Oxide Nanoparticles (IONPs) have been synthesized under different rates of addition of the base NaOH. The efficiency of SIONPs and IONPs for removal of hexavalent chromium from aqueous medium has been studied. The variation of the adsorption capacity with pH depends on the nanoparticle phase composition as well as its starch functionalization. Cr(VI) removal by SIONPs and IONPs follows a Langmuir adsorption isotherm. The highest monolayer saturation adsorption capacity as obtained from the Langmuir adsorption isotherm for SIONPs is 9.02 mg g−1.
Sodium carboxymethyl cellulose and starch coated with nanoscale zerovalent iron-nickel (SS–nZVI-Ni) were devised and their Cr(VI) removal extent was analyzed and crosschecked.190 The results showed that SS–nZVI-Ni has high Cr(VI) removal aptitude. Cr(VI) removal by SS–nzvi-Ni while being affected by acidic conditions revealed that the Cr(VI) removal efficiency by SS–nZVI-Ni reached a maximum of 95.70% at pH 2. Different initial Cr(VI) concentrations showed that SS–nZVI-Ni performed well at a high Cr(VI) concentration. Meanwhile, SEM images showed that SS–nZVI-Ni maintained a larger surface area, solving aggregation problems.
Greenly synthesized alginate nanoparticles stabilized by honey were prepared by cross-linking an aqueous solution of alginate with calcium ions. The potential of applying these nanoparticles for removing Cr(VI) was studied, where a maximum efficiency of removal of 93.5% was obtained.192 The method reported in that study affords the synthesis of many different protein and polysaccharide nanoparticles like chitosan, casein, and albumin. Since the materials used are all nontoxic and biocompatible, this method can also be used in the synthesis of nanoparticles for biomedical applications. Alginate nanoparticles synthesized using honey were highly stable. Even though they showed slight aggregation on drying, they can be redispersed in water by mild sonication. Various microscopic analyses confirmed the formation of stable nanoparticles. The presence of honey prevented the stickiness and aggregation of nanoalginate particles on centrifugation. The high adsorption capacity of the synthesized nanoparticles explores the possibility of removing toxic metal ions from contaminated water.
Sodium alginate was used to encapsulate magnetite graphene oxide to create mGO/alginate beads. Cr(VI) and As(V) adsorption by the mGO/alginate beads agreed with pseudo-second-order kinetics.155 The overall ranking for fit isotherm models is Freundlich > Langmuir for Cr(VI) and Langmuir > Freundlich for As(V). Core material (mGO) aggregation and leaching were avoided; thus, the adsorption capacity was enhanced compared with other adsorbent materials. By applying various cross-linking cations, the pH at the interface could be controlled at the optimal conditions for removing Cr(VI) and As(V). The mGO/alginate beads possess the potential to optimally displace heavy metals from a multicomponent system and contaminated wastewater. Moreover, mGO/alginate beads could be easily separated from aqueous solutions due to their superparamagnetic nature, thus avoiding secondary pollution. With all the advantages mentioned beforehand, the mGO/alginate beads are anticipated to be a diverse material in the treatment of water.
Auricularia auricula spent substrate altered by cetyltrimethyl ammonium bromide (CTAB) and stabilized on sodium alginate was tested as a novel biosorbent (MIAASS) for removing Cr(VI) in a fixed-bed column.193 The Thomas model was more reasonably accurate in predicting the experimental column capacity than the Adams–Bohart model. Following three cycles of the adsorption–desorption process, the regeneration efficiency of MIAASS was 52.48%.
Omer et al. (2019)194 developed an efficient adsorbent based on tetraethylenepentamine (TEPA)-functionalized alginate (Alg) beads (TEPA-Alg) for the adsorptive removal of Cr(VI) from aqueous solutions. The alginate surface's chemical modification was performed via p-benzoquinone (PBQ) as a coupling agent followed by functionalization with TEPA. The concentration of amine groups in TEPA-Alg surged with the increase of the TEPA concentration up to 0.05 M. The functionalized beads illustrated a higher affinity towards the adsorption of Cr(VI) compared with native alginate beads. The maximum Cr(VI) removal percentage was 100% and was obtained using 0.2 g of adsorbent, contact time 180 min, pH 2, room temperature, and 50 rpm. Moreover, the experimental data were best fit to the pseudo-first-order model while the adsorption equilibrium adhered to a Langmuir adsorption isotherm, i.e., maximum adsorption capacity ∼77 mg g−1. Additionally, the study of reusability showed good adsorption traits after five cycles in a row. Therefore, the prepared functionalized TEPA-Alg beads could be optimally used for removing Cr(VI) ions from their aqueous solutions.
Another type of chemically modified alginate beads for removing heavy metals were synthesized via polyethylenimine (PEI) as a modification reagent, with high-density active sites introduced favorably onto the surface by a cross-linking reaction; this led to designing a flexible core–shell/bead-like alginate@PEI sorbent which was employed for Cr(VI) adsorption from aqueous solutions.195 The maximum adsorption extent of Cr(VI) on the typical alginate@PEI-1.5 was 431.6 mg g−1, surpassing most reported sorbents for Cr(VI) removal from aqueous solutions. The adsorption kinetics followed the pseudo-second-order kinetic equation while the Freundlich model optimally described the equilibrium data of adsorption of Cr(VI). The alginate@PEI adsorption capacity only had a slight loss after ten cycles of adsorption–desorption studies. Compared with other reported adsorptive materials, this core–shell/bead-like composite promises advantages of low cost, ultra-high adsorption capacity, environmental friendliness, and exemplary recycling performance.
Alginate can be made into various usable forms like membranes, beads, and candles for water remediation. To prepare alginate in a usable bead form, magnetic nano-hydroxyapatite encapsulated alginate beads (Fe3O4@n-HApAlg) were devised by a hydrothermal method for selective removal of chromium.196 The thermodynamic studies imply the spontaneous and endothermic nature of the sorption of Cr(VI). Fe3O4@n-HApAlg beads were made usable by using NaOH eluent, up to five cycles. The magnetic alginate hybrid beads showed promising results against chromium-contaminated groundwater by removing Cr(VI) ions below the tolerance limit. Regeneration and reusability assessments were successfully conducted, and the hybrid beads can be reused for up to five successive cycles. The synthesized Fe3O4@n-HApAlg beads were an optimal material for the treatment of Cr(VI)-contaminated field water.
Removing Cr(VI) using NZVI immobilized in alginate beads (NZVI beads) and a bacterial consortium immobilized in alginate beads (bio beads) in a sequence process was performed.197 The process utilizing the NZVI beads and bio beads was successfully analyzed with a Cr(VI)-spiked groundwater sample. The efficiency of removing Cr(VI) was observed to upscale (more than 90 ± 5.96% at 381 ± 17.1 mg g−1 capacity for a 10 mg L−1 influent concentration) using NZVI beads and microbial consortium immobilized beads in sequence. The system was successfully assessed with real water systems, having direct field applicability for that process.
The suitability of the Ca-alginate immobilized consortium SFC 500-1 for simultaneous removal of Cr(VI) and phenol was demonstrated.198 Immobilization was a successful strategy to enhance the removal of Cr(VI), mostly associated with enzymatic reduction through the synthesis of Cr(III). The immobilized SFC 500-1 showed great prowess in degrading phenol in a short time, probably due to the protective impact of the alginate matrix against the toxicity of phenol concentrations over 1000 mg L−1. Complete phenol oxidation to catechol and cis-muconate was verified through the detection of these intermediates in the reaction medium during the co-remediation process. The optimum long-term storage, removal potential, and reusability make SFC 500-1 entrapped cells an interesting system for the removal of Cr(VI) and phenol in media of low-concentration organic matter. The low-cost and high-resistance alginate matrix exhibits more merits when possible applications for the decontamination of real effluents are considered.
Ali (2018)200 prepared chitosan-1,2-cyclohexylenedinitrilotetraacetic acid/graphene oxide (chitosan/CDTA/GO) nanocomposite in the presence of glutaraldehyde as a cross-linker. The removal efficiency and adsorption of Cr(VI) utilizing the chitosan/CDTA/GO adsorbent were investigated at various adsorption conditions. The prepared adsorbent at hand was characterized by SEM and FTIR. The effects of the adsorbent chemical composition, CDTA/GO concentration, adsorbent dose, pH, temperature, time of contact, and initial metal ion concentration on Cr(VI) sorption were investigated. The results showed that the optimum adsorbent dose was 2 g L−1 at a pH value of 3.5 and an equilibrium time of 60 min. Besides, the temperature and pH immensely influenced the adsorption process. The Cr(VI) adsorption kinetics onto Cs/CDTA/GO followed the pseudo-second-order model and the adsorption isotherm adhered to the Langmuir model. The maximum adsorption capacity of the adsorbent was 166.98 mg g−1, and the equilibrium parameter (RL) at different concentrations was less than unity, indicating that Cr(VI) ion adsorption onto chitosan/CDTA/GO is favorable. The chitosan/CDTA/GO adsorbent could be regenerated more than three times according to its adsorption–desorption cycles.
Chitosan and alginate nanocomposites were evaluated for chromium(VI) removal from wastewater.201 The sorbent was characterized by XRD, FT-IR, SEM, DLS, and DSC. Batch adsorption experiments were carried out to evaluate the removal process under various factors like the effects of the initial concentration, dose of adsorbent, pH value, and agitation time. The pH-dependent metal ion removal reached the optimum at pH 5.0. The experimental data were analyzed by the Freundlich and Langmuir isotherms. The isotherm study revealed that the adsorption equilibrium is adherent to the Freundlich isotherm and the sorption extent of alginate and chitosan nanocomposites is very high, and the adsorbent favors multilayer adsorption. Pseudo-first and second-order kinetics models were used to illustrate the kinetic data. It was determined that removing Cr(VI) adhered to second-order reaction kinetics. It is concluded that alginate and chitosan nanocomposites are exemplary biosorbents material for adsorping Cr(VI) from waterwaste.
Zirconium(IV) cross-linked chitosan magnetic microspheres (Fe3O4@Zr-chitosan) as a recoverable adsorbent were fabricated through the coordination reaction between zirconium oxychloride and a chitosan biopolymeric matrix for efficient adsorption and simultaneous detoxification of Cr(VI) in aqueous solutions.202 XRD and SEM verified the formation of core@shell magnetite microspheres. XPS and FT-IR confirmed Zr(IV)'s cross-linking with chitosan on the microspheres. The batch Cr(VI) adsorption performances of the resultant Fe3O4@Zr-chitosan microspheres showed that a maximum adsorption capacity of 280.97 mg g−1 was achieved under pH 4.0 at 298 K. The XPS analyses indicated that 61.1% of the adsorbed Cr(VI) was reduced to Cr(III) because of the oxidization of alcoholic groups on C-6 in chitosan serving as electron donors to carbonyl groups. The adsorbent showed preferential Cr(VI) adsorption with the existence of coexisting cations (K+, Na+, Cu2+, Zn2+, Ca2+, and Mg2+) and anions (NO3−, Cl−, SO4−2, and CO3−2). The adsorbent maintained excellent reusability and lowered the effluent Cr(VI) contents down to the ppb level, reaching the drinking water standard recommended by the World Health Organization, and was a potential option for water purification.
Chitosan–Fe(OH)3 beads were one-step synthesized without using acid solvents, which could be used as effective adsorbents for anionic dye removal. The preparation process was simple and green.203 The prepared beads were characterized by morphological and structural analyses in detail using several techniques, e.g., SRD, SEM, XRD, TGA, and XPS. The Fe(OH)3 content in the chitosan–Fe(OH)3 beads was 54.64 wt%. The removal efficiencies towards anionic dyes, Congo Red (CR) and Methyl Orange (MO), by the chitosan–Fe(OH)3 beads surpassed the pure chitosan beads. Moreover, incorporating Fe(OH)3 into chitosan beads could overcome the shortcoming that powdery Fe(OH)3 particles are not easy to separate from adsorption solutions. The maximum adsorption capacities from the Langmuir model for CR and MO by the chitosan–Fe(OH)3 beads were 445.32 and 314.45 mg g−1, respectively. The calculated thermodynamic parameters, namely ΔG°, ΔH°, and ΔS°, suggest that adsorption of CR and MO was endothermic and spontaneous in the temperature range 293–353 K. More importantly, the chitosan–Fe(OH)3 beads can be regenerated and the adsorption properties reset after NaOH treatment. The desorption results showed that the adsorption capacity can remain up to 95% after being used five times. Moreover, the chitosan–Fe(OH)3 beads additionally exhibited good reusability and the efficiencies for the removal of both dyes surpassed 95% after five cycles. From this work, it is suggested that chitosan–Fe(OH)3 beads have great potential as effective and low-cost adsorbents for removing anionic dyes.
Bimetallic Fe/Cu nanoparticles were successfully stabilized by chitosan and then used for remediating Cr(VI)-contaminated wasterwater.204 However, the overloaded chitosan on the surface of Fe/Cu particles limited the reduction of Cr(VI) due to the occupation of the surface reactive sites. Most importantly, the contribution of chitosan and Cu in the removal mechanism was studied by reduction experiments and XPS analysis. On one hand, chitosan could effectively combine with Cr(VI) due to chelation; on the other hand, Cu played an important role in the precipitation and coprecipitation phenomena. These findings dictate that CS–Fe/Cu has the potential to be a promising material for wastewater treatment.
The preparation of Cu2S–chitosan nanocomposites for removing Cr(VI) ions from wastewater samples was achieved.205 Structural studies of the particles were carried out using TEM and XRD. The XRD patterns presented the cubic Cu1.8S phase for all the synthesized copper sulfide nanoparticles. TEM images showed spherical shape particles which became agglomerated when the stabilizers were introduced into the nanomaterial. The adsorption studies proved that the pH value, dosage, and time of contact play an important part during the water treatment.
Guo et al. (2018)206 synthesized polydopamine-modified-chitosan (CS-PDA) aerogels through dopamine self-polymerization and glutaraldehyde cross-linking reactions to increase the capacity of adsorption and acid resistance of chitosan. CS-PDA were applied in the adsorption of Cr(VI) and Pb(II). CS-PDA exhibited superior adsorption performances in removing Cr(VI) and Pb(II). The adsorption isotherms and kinetic data were adherent to the Langmuir and pseudo-second-order kinetic models. The maximum adsorption capacities of CS-PDA for Cr(VI) and Pb(II) were 374.4 and 441.2 mg g−1, respectively. After eight cycles, the CS-PDA adsorption capacity showed no obvious decline. These superiorities make CS-PDA a promising multifunctional adsorbent for the purification of metal ions and dyes.
A chitosan-grafted graphene oxide (CS–GO) nanocomposite was tested for the adsorption of Cr(VI) in batch mode.207 The CS–GO nanocomposite material was prepared by the ultrasonic irradiation technique. The CS–GO adsorbent was characterized by XRD, FTIR spectroscopy, SEM, and TEM. An adsorption capacity of 104.16 mg g−1 was achieved at pH 2.0, at a contact time of 420 min. The adsorption process was confirmed by the pseudo-second-order kinetic and Langmuir isotherm models. The nano-microstructural investigation validates the successful Cr(VI) adsorption on the CS–GO nanocomposite. The CS–GO material is recyclable for up to ten cycles with a minimum loss in adsorption capacity.
Thiosemicarbazide-modified chitosan (TSFCS) was fabricated and characterized by SEM, FTIR, and TGA techniques and its application as an adsorbent for the displacement of Pb(II) and methyl red (MR) was evaluated.208 Some adsorption parameters, e.g., the dose of sorbent, pH value, and initial concentration of solutions, were evaluated alongside their effects. The maximum removal percentages for MR and Pb(II) were found to be about 91% and 85.6% at pH values 8 and 5 and sorbent dosages of 3.33 g L−1 and 1.33 g L−1, respectively. Langmuir, Freundlich, and Temkin isotherms were evaluated and the maximum adsorption extents for Pb(II) and MR were found to be 56.89 mg g−1 and 17.31 mg g−1 at 55 °C and 25 °C, respectively. Isotherm studies indicate that Pb(II) and MR adsorption adhere to the Freundlich and Langmuir models, respectively. The kinetics of adsorption was more accurately adherent to the PSORE. Thermodynamic parameters including ΔG°, ΔH°, and ΔS° were gauged over the temperature range of 25–55 °C. The sorption process was endothermic and spontaneous.
Neto et al. (2019)209 prepared an adsorbent in the form of an iron oxide/carbon nanotube/chitosan magnetic composite film (CLCh/MWCNT/Fe). CLCh/MWCNT/Fe was characterized by Inductively Coupled Plasma Mass Spectrometry (ICP-MS), nitrogen adsorption/desorption, SEM, XRD, Energy-Dispersive Spectrometry (EDS), Raman spectroscopy, and IR. The CLCh/MWCNT/Fe film presented a maximum adsorption capacity of 449.30 mg g−1 for Cr(VI) (60 min) at 25 °C. In ten consecutive reutilization adsorption cycles, the CLCh/MWCNT/Fe film suffered efficiency losses of only 12% and 6% for the removal of Cr(III) and Cr(VI), respectively.
C 0 | Sorbate initial concentration (mol L−1) |
C e | The equilibrium concentration (mol L−1) |
q e | The amount of sorbate sorbed in the absorbent at equilibrium (mol g−1) |
q m,L | The capacity of maximum monolayer (mol g−1) |
K L | Langmuir isotherm constant (L mol−1) |
K F | Freundlich constant relative sorption capacity (mol g−1) (L mol−1)1/n |
n | Sorption intensity |
R | Universal gas constant (8.314 J mol−1 K−1) |
T | Temperature (K) |
A T | Temkin isotherm equilibrium constant (L g−1) |
b T | Temkin isotherm constant (J mol−1) |
ε | The Polanyi potential (J2 mol−2) |
q s | Theoretical isotherm saturation capacity (mol g−1) |
β | Dubinin–Radushkevich isotherm constant (mol2 kJ−2) |
θ | Degree of surface coverage |
K FH | Flory–Huggins isotherm equilibrium constant (L g−1) |
K D | Hill constant |
n H | Hill cooperativity coefficient of the binding interaction |
q sH | Hill isotherm maximum uptake saturation (mol L−1) |
K R | Redlich–Peterson isotherm constant (L g−1) |
a R | Redlich–Peterson isotherm constant (L mol−1) |
g | Redlich–Peterson isotherm exponent |
K S | Sips isotherm model constant (L mg−1) |
β S | Sips isotherm model exponent |
K T | Toth isotherm constant (mol g−1) |
B | Koble–Corrigan isotherm constant (L mol−1)n |
A | Koble–Corrigan isotherm constant (Ln mol1−n g−1) |
b k | Khan isotherm model constant |
α k | Khan isotherm model exponent |
α RP | Radke–Prausnitz isotherm model constant |
β R | Radke–Prausnitz isotherm model exponent |
γ R | Radke–Prausnitz isotherm model constant |
α | Frenkel–Halsey–Hill isotherm constant (J mol−1) |
r | Sign of inverse power of distance from the surface |
d | Interlayer spacing (m) |
k | MacMillan–Teller (MET) isotherm constant |
q t | Amount of metal ions sorbed at time t (mmol g−1) |
K 1 | Pseudo-first-order rate constant of adsorption (min−1) |
K 2 | Pseudo-second-order rate constant of adsorption (g mmol−1 min−1) |
K i | The intra-particle diffusion rate (mmol g−1 min−0.5) |
X | The boundary layer diffusion effects (mmol g−1) |
α | The initial sorption rate (mmol g−1 min−1) |
β | The desorption constant (g mmol−1) |
This journal is © The Royal Society of Chemistry 2020 |