Next Article in Journal
Effects of Hybrid-Type Artificial Groundwater Recharge and Underground Barrier in a Small Basin
Previous Article in Journal
Synergistic Effects and Ecological Responses of Combined In Situ Passivation and Macrophytes toward the Water Quality of a Macrophytes-Dominated Eutrophic Lake
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Dynamic Adsorption of As(V) onto the Porous α-Fe2O3/Fe3O4/C Composite Prepared with Bamboo Bio-Template

1
College of Environmental Science and Engineering, Guilin University of Technology, Guilin 541004, China
2
Guangxi Key Laboratory of Environmental Pollution Control Theory and Technology, Guilin University of Technology, Guilin 541004, China
3
Collaborative Innovation Center for Water Pollution Control and Water Safety in Karst Area, Guilin University of Technology, Guilin 541004, China
*
Author to whom correspondence should be addressed.
Water 2022, 14(12), 1848; https://doi.org/10.3390/w14121848
Submission received: 26 April 2022 / Revised: 1 June 2022 / Accepted: 6 June 2022 / Published: 8 June 2022
(This article belongs to the Section Wastewater Treatment and Reuse)

Abstract

:
Arsenic (As(V)), a highly toxic metalloid, is known to contaminate wastewater and groundwater and is difficult to degrade in nature. However, the development of highly efficient adsorbents, at a low cost for use in practical applications, remains highly challenging. Thus, to investigate the As(V) adsorption mechanism, a novel porous α-Fe2O3/Fe3O4/C composite (PC-Fe/C-B) was prepared, using bamboo side shoots as a bio-template, and the breakthrough performance of the PC-Fe/C-B composite-packed fixed-bed column in As(V) removal was evaluated, using simulated wastewater. The PC-Fe/C-B material accurately retained the hierarchical porous microstructure of the bamboo bio-templates, and the results demonstrated the great potential of PC-Fe/C-B composite, as an effective adsorbent for removing As(V) from wastewater, under the optimal experimental conditions of: influent flow 5.136 mL/min, pH 3, As(V) concentration 20 mg/L, adsorbent particle size < 0.149 mm, adsorption temperature 35 °C, PC-Fe/C-B dose 0.5 g, and breakthrough time 50 min (184 BV), with qe,exp of 21.0 mg/g in the fixed-bed-column system. The CD-MUSIC model was effectively coupled with the transport model, using PHREEQC software, to simulate the reactive transportation of As(V) in the fixed-bed column and to predict the breakthrough curve for column adsorption.

1. Introduction

The contamination of wastewater and groundwater with arsenate (As(V)) is one of the most concerning environmental issues globally [1,2]. Arsenic is a highly toxic metalloid that is released into the environment, mainly through human activities. It has heavy metal characteristics, in that, it is difficult to degrade in nature and undergoes bioaccumulation; it can enter the human body through the food chain, drinking water, respiration, and other ways, thereby posing serious dangers to human health. It is, also, listed as a group 1 carcinogen because of its high toxicity and carcinogenicity [3,4]. However, it, also, has certain non-metal characteristics. For example, As-containing compounds are unstable in water, readily decompose to amphoteric oxides, and easily migrate and transform in the environment [5,6].
In recent years, treatment of wastewater containing arsenic has received increasing attention. In this respect, Glass [7] modified polyacrylonitrile ultrafiltration membranes with amine functionalities, and the modified membrane showed that the arsenate adsorption capacity was five times higher than that of a pristine PAN membrane. Lee [8] developed a new beaded adsorbent, using coal-mine-drainage sludge, and this effectively removed arsenic and other heavy metals from acid-mine drainage. Of all the As(V)-removal technologies, adsorption has been found to be the simplest and most effective method [9,10,11]. Adsorption has multiple advantages for use in water decontamination, including its low cost, high efficiency, and ease of operation as well as the absence of secondary-pollutant production. However, developing highly efficient adsorbents, at a low cost with good adaptability to packed-bed operation, for use in practical applications, remains highly challenging.
Previous studies have shown that effective As(V) adsorbents include iron oxide, aluminium oxide, clay, rare-earth oxides, and coal-fly ash. Among these, iron oxides (such as haematite, magnetite, goethite, akaganeite, feroxyhyte, siderite, and iron hydroxide) have been proven to be the best As(V) adsorbents [12,13,14,15,16,17]. Iron oxides can effectively immobilise As(V) through sorption processes because of their large surface areas and charged reactive surface sites, which provide extremely high sorption capacities [18,19]. By taking advantage of iron oxides and the hierarchical porous structure of Moso bamboo, a porous α-Fe2O3/Fe3O4/C composite (PC-Fe/C-B) was prepared, using the side shoots of Moso bamboo as a microstructural template [20].
The surface-complexation model (SCM) has been employed to study the adsorption behaviour of As(V) on PC-Fe/C-B, and it is a vital tool used in studying the adsorption behaviour of ions. The charge-distribution-multi-site complexation (CD-MUSIC) model [21] is an advanced SCM for ion adsorption to metal oxides [22], and it is one of the most widely used models to describe As adsorption behaviour on metal-oxide surfaces. The CD-MUSIC model simultaneously considers the heterogeneity of the adsorption sites on the surface of the adsorbent, the spatial distribution of the charge of the adsorption centre atom in the electric double layer, and the contribution to the charge neutralisation of the oxygen atom in the adjacent electric layer. It, also, follows the three core principles of mass action, mass balance, and electrostatic interactions between reactants. A surface reaction equation system is constructed, and a chemical-equilibrium calculation is conducted under certain constraints. The simulation results can be used to predict the possible surface-complex forms after adsorption, such as inner or outer adsorption, monodentate or pluridentate surface complexes, protonation and protonation degree, and surface-charge change. The CD-MUSIC model is, mainly, based on actual experimental data, such as adsorption isotherms and pH, which are increasingly used in the study of adsorption processes at solid/liquid interfaces. In this respect, Salazar-Camacho and Villalobos (2010) [23] modelled arsenate adsorption on goethite, using experimental data from several previously published studies. Gustafsson (2001) [24] studied the adsorption of As(V) in soil, using the CD-MUSIC model, and found that oxide minerals (ferrihydrite and gibbsite) and clay minerals (proto-imogolite allophane) in the soil were the main adsorption interfaces. Furthermore, Liang et al. (2021) [25] revealed the complex species of Pb on the haematite surface, using CD-MUSIC modelling.
In this study, the PC-Fe/C-B adsorbent was prepared, using bamboo side shoots as a bio-template, which was systematically characterised using scanning-electron microscopy (SEM), energy-dispersive X-ray spectroscopy (EDS), Brunauer–Emmett–Teller (BET) surface-area analysis, X-ray diffraction (XRD), X-ray photoelectron spectroscopy (XPS), and Fourier-transform infrared (FT-IR) spectroscopy. The adsorption performance of PC-Fe/C-B was, also, evaluated, using a fixed-bed-column system with simulated As(V)-containing wastewater, and the adsorption kinetics, influencing factors, and adsorption mechanism of As(V) on the composites were systematically investigated. The adsorption data were, subsequently, analysed in detail, using several models, including the Thomas model, the Yoon–Nelson model, the Adams–Bohart model, the Clark model, and the Wolborska model. For further study of adsorption behaviour, CD-MUSIC modelling was conducted, using PHREEQC software (Version 3.0), which integrates CD-MUSIC with the transport block. PHREEQC [26] is a computer program used to simulate geochemical reactions and migration processes in natural or polluted water, and it performs complex speciation, intermittent reactions, one-dimensional (1D) transport, and geochemical-inversion simulations. As such, it has been widely applied in thermodynamic studies. Based on the characterisation of the PC-Fe/C-B surface and the CD-MUSIC simulation, with the program PHREEQC, the speciation distribution and surface complexation on the PC-Fe/C-B adsorbents, during As(V) adsorption, were analysed to investigate the adsorption mechanism.

2. Materials and Methods

2.1. Preparation and Characterization

2.1.1. Materials and Reagents

Moso bamboo was obtained from Guangxi, China. The reagents Fe (NO3)3, ethanol, Na3AsO4·12H2O, HNO3, NaOH, and 5% (v/v) ammonia solutions were purchased from the Country Medicine Group, Shanghai, China. All reagent solutions were prepared from analytical-grade chemical reagents and ultrapure water.

2.1.2. Sample Preparation

The side shoots of Moso bamboo were pre-treated, according to the method presented by Zhu et al. (2013) [20], and used as a template for preparing the composite adsorbent. In brief, the adsorbent PC-Fe/C-B was prepared by extraction pre-treatment, cleaning, drying, cyclic soaking in ferric nitrate solution, and roasting. Specifically, the following procedure was followed: (1) Chopped bamboo shoots were heated in 5% dilute ammonia, at 100 °C for 6 h, followed by washing with ultrapure water and drying at 80 °C for 24 h. (2) The pre-treated bamboo templates were immersed in a 1.2 mol/L precursor solution, at 60 °C, which was prepared by dissolving ferric nitrate in a mixture of ethanol and ultrapure water (1:1 volume ratio). The bamboo templates were soaked in the solution for 5 d, and the precursor solution was regularly checked to ensure even immersion. Subsequently, the samples were removed from the solution and oven-dried, at 85 °C for 1 d. The soak-dry process was repeated thrice. (3) The samples were carbonised in a muffle furnace, by slow heating to 600 °C and holding at this temperature for 3 h. After cooling to room temperature, samples with different particle sizes were obtained, by grinding and screening.

2.1.3. Characterization of PC-Fe/C-B Sorbent

The specific surface area and porosity distribution of PC-Fe/C-B sorbent with different particle sizes ranging from 1.7–300.0 nm were measured by Quantachrome NOVAe1000, using a nitrogen adsorption–desorption method, with helium as the carrier gas and nitrogen as the adsorbate, at a liquid-nitrogen temperature of 77 K. All samples were degassed at 150 °C, in vacuum. The porosity of large pores (pore sizes of 0.0071–0.60 μm) was tested, using a mercury porosimeter (PoreMasterGT 60, Quantachrome, Boynton Beach, FL, USA). The test conditions were as follows: sample volume, 0.0977 g; mercury surface tension, 480 erg/cm2; and mercury contact angle, 140°. The surface morphology of the PC-Fe/C-B sorbent was examined using SEM (JEOL JSM-6380LV, Japan Electron Optics Ltd., Mitaka, Tokyo), and its chemical composition was assessed using EDS. The as-prepared samples were characterised by XRD analysis (X’Pert PRO, PANalytical B.V., the Netherlands), using Cu Kα radiation (λ = 0.15406 nm). The samples were, also, measured in the form of KBr pellets, over the range 4000–400 cm−1, using FT-IR spectroscopy (Nicolet Nexus 470, Thermo Fisher Scientific Inc.). XPS measurements, using an X-ray photoelectron spectrometer (VG ESCALab Mark II spectrometer; VG Scientific Ltd., East Grinstead, UK), were conducted on samples before and after adsorption.

2.2. Arsenate Adsorption to PC-Fe/C-B

Simulated wastewater containing As(V) was prepared, using Na3AsO4·12H2O and HNO3. Dynamic-adsorption experiments were performed, using a small fixed-bed-column adsorption device. The fixed-bed column had a length of 200 mm, with outer and inner diameters of 35 mm and 12 mm, respectively. The concentration of As(V) in the aqueous solution was analysed, by an AAS flame and graphite furnace atomic absorption spectrophotometer (AA700, Perkin Elmer, Boston, MA, USA).
The factors influencing the dynamic adsorption of As(V) were investigated, by changing the initial As(V) concentration of the influent (10 mg/L, 20 mg/L, 30 mg/L, and 50 mg/L), influent flow rate (4.28 mL/min, 5.136 mL/min, 6.849 mL/min, and 8.561 mL/min), influent pH (1.0, 2.0, 3.0, 4.0, 5.0, 6.0, and 7.0), adsorbent dose (0.3 g, 0.4 g, 0.5 g, and 0.6 g), particle size of the adsorbent (0.841–0.4 mm, 0.4–0.25 mm, 0.25–0.177 mm, 0.177–0.149 mm, and <0.149 mm), and adsorption temperature (25 °C, 35 °C, and 45 °C). To obtain the main dynamic adsorption-process parameters and to determine the best model to describe the dynamic adsorption behaviour, the adsorption behaviour was analysed, using several models, including the Thomas model [27], the Nelson model [28,29], the Bohart model [30], the Clark model [31], and the Wolborska model [32].

2.3. Modeling

The CD-MUSIC model is based on the SCM approach, developed by Vaughan Jr. et al. [33,34]. It can be used to describe the adsorbent surface-acid–base reactions and adsorbate–surface reactions that constitute chemical equilibrium, and it was, thus, selected to model the surface adsorption behaviour in this study. Simulations and parameter optimisation were performed, using PHREEQC (Version 3.0). The complete CD-MUSIC model incorporates a basic Stern-layer approach, iron-oxide-surface characteristics, As(V)-aqueous-phase speciation, surface-acid–base reactions, and adsorbate–surface reactions.
Two data blocks of PHREEQC were used to achieve the coupled simulations for As(V) migration–adsorption [26], which were advection and transport. In the transport data block of this paper, an adsorption column composed of 10 cells was simulated with a total amount of adsorbent weighing 0.5 g = (10 × 0.05 g), total adsorption column height of 1.23 cm = (10 × 0.123 cm), and inflow rate of Q = 5.136 mL/min. The water-chemistry parameters, PC-Fe/C-B-surface properties, aqueous-phase As speciation, transport-parameter modelling, and surface complexation are described in detail in the PHREEQC input file and the explanation notes in the electronic supplementary material (Tables S6–S9).

3. Results and Discussion

3.1. Characterization of PC-Fe/C-B

The BET method was used to determine the specific surface area of PC-Fe/C-B, with different particle sizes at 77 K. The BET-specific surface areas were determined as ranging from 138.17 m2/g to 186.66 m2/g, with corresponding pore volumes ranging from 0.07487 cm3/g to 0.11938 cm3/g, which suggests that PC-Fe/C-B exhibited good adsorption potential. Under the test range of pore sizes ranging from 1–300 nm, most of the PC-Fe/C-B pores ranged from 2.0–19.0 nm (74–81.3%)—within the mesopore range, although some were within the micropore range of <2 nm (18.3–26%). The results indicated that PC-Fe/C-B was a hierarchically structured porous material with a medium pore size. The large pore size and porosity of blocky PC-Fe/C-B were determined, using the mercury injection method. PC-Fe/C-B was found to contain abundant pores and to have a porosity of 69.51%.
The SEM results, also, showed that PC-Fe/C-B preserved the hierarchical porous structure of the bamboo wood (Figure 1). Three different types of pores (macropores, mesopores, and micropores) were retained from the bio-templates, and these originated from vessels (widths 50–70 μm), parenchyma cells (10–30 μm), fibre pores (widths 3–10 μm) and pits on the walls of the vessels and parenchyma cells (widths 0.1–1.3 μm), respectively. It is known that interpore connectivity can be significantly increased by extraction, pre-treatment, and high-temperature calcination. Quantitative SEM–EDS peak-area analysis showed that up to 4.19% (w/w) As(V) was adsorbed on the PC-Fe/C-B material (Figure 1). EDS spectra were collected, from samples treated with an As(V) concentration of 100 mg/L.
The XRD results from samples, before and after As(V) adsorption, are shown in Figure 2, indicating that magnetite, hematite, and carbon were present on the PC-Fe/C-B surface. In the preparation process, the cellulose in the bamboo template went through the processes of dehydrogenation and deoxidation by high-temperature calcination, and the adsorbed-ferric-nitrate precursor chemically decomposed into iron oxide (α-Fe2O3). In a high-temperature oxygen-containing environment, part of the iron oxide obtained from decomposition was further transformed into magnetic iron oxide (Fe3O4). Therefore, the crystal forms of iron oxide in PC-Fe/C-B were Fe3O4 and α-Fe2O3. It can be seen from Figure 2 that the characteristic peaks of α-Fe2O3, at 2θ = 24.50°, 40.85°, 49.53°, and 63.98°, disappeared after As(V) adsorption; the main characteristic peaks for α-Fe2O3 at 2θ of 33.21° and 54.12° were significantly weakened; and the characteristic peaks for Fe3O4 at 2θ of 29.98°, 43.34°, 57.50°, and 62.42° were strengthened, which indicated that the crystal form of iron oxide changed after adsorption. It was speculated that this may be due to the hydration reaction of some α-Fe2O3 dissolved in water, to generate FeOOH or a surface-complexation reaction.
In the FT-IR spectra (Figure 3) of PC-Fe/C-B after As(V) adsorption, the presence of a wide absorption peak at 3415.26 cm−1 was attributed to the contracting vibration of −OH in the material, rather than to the dissociation of physically adsorbed water (due to the absence of a vibration peak near 1630 cm−1) [35]. This, also, showed that the number of −OH groups on the surface of iron oxide increased after As(V) adsorption, which, also, confirmed the acid–base dissociation behaviour of PC-Fe/C-B, during the adsorption of As. After As adsorption, the intensity of the stretching vibration peak of CO2 at 2361 cm−1 decreased, and the C-O and C=O vibration peaks [35], at 1541 cm−1 and 1516 cm−1, shifted to higher wave numbers, of 1551 cm−1 and 1532 cm−1, respectively. The Fe-O vibration peaks, at 564 cm−1 and 443 cm−1, also, shifted to lower wavenumbers, of 556 cm−1 and 440 cm−1, indicating that the groups corresponding to these peaks were the active adsorption sites on the surface of PC-Fe/C-B. In the FT-IR spectra of the PC-Fe/C-B sorbent after sorption of As(V), peaks of AsO43− for As–O were observed to be bending around 873.60 cm−1.
The XPS survey spectra and high-resolution Fe 2p, C 1s, and O 1s spectra of PC-Fe/C-B, after As(V) adsorption, are shown in Figure 4. After As(V) adsorption, the As 3d peak, with a binding energy of 42.78 eV, was detected, which indicated that As existed as As(V) without a redox reaction [36]. The O 1s and Fe 2p peaks, after As(V) adsorption by PC-Fe/C-B, showed almost no shift, indicating that the chemical states of O and Fe did not change after As(V) adsorption. The atomic percentages of PC-Fe/C-B, after As adsorption, were 58.96% C, 28.72% O, 11.07% Fe, and 1.25% As. Moreover, since the charge amount of the above electron transfer was far less than a unit of electricity, it was speculated that the transfer of the outer shared electron pair occurred during the bonding process.

3.2. As(V) Dynamic Adsorption

The results showed that the PC-Fe/C-B adsorbent could be used to remove As(V) effectively from wastewater, under the following optimal experimental conditions: influent flow 5.136 mL/min, pH 3, As(V) concentration 20 mg/L, adsorbent particle size < 0.149 mm, adsorption temperature 35 °C, PC-Fe/C-B dose 0.5 g, and breakthrough time 50 min (184 BV), with qe,exp of 21.0 mg/g. Increasing the PC-Fe/C-B adsorbent amount (adsorption-column height) and reducing the particle size, effectively prolonged both the breakthrough time and failure time as well as improved the quality of the column, thus improving the performance of the adsorption column. In addition, the total As(V) removal rate increased with increasing PC-Fe/C-B dose. However, increasing the influent concentration and flow rate accelerated column failure. PH had a significant influence on As(V) adsorption, and acidic conditions were conducive to adsorption (pH 2–4). Temperature had little effect on the adsorption of As(V) by PC-Fe/C-B. The effects of each factor on As(V) adsorption are shown in Figure 5, Figure 6, Figure 7, Figure 8, Figure 9 and Figure 10.

3.3. As(V) Adsorption Capacity of PC-Fe/C-B

Table S5 in the electronic Supplementary Material compares the As(V) adsorption capacities of PC-Fe/C-B with various adsorbents reported in the literature. As shown in Table S5, PC-Fe/C-B has a great potential for the adsorption and removal of As(V) from solution. PC-Fe/C-B has a larger adsorption capacity than several adsorbents tested under similar conditions that investigated particle size and initial concentration. Some of these adsorbents are as follows: an aluminium mining by-product (4.51 mg/g) [37], PBGC-Fe/C (4.83 mg/g) [20], organo-modified natural zeolite materials (6.7 mg/g) [38], biochar-magnetite composite (5.49 mg/g) [14], Chinese red soil (0.936 mg/g) [39], iron-coated seaweeds (7.3 mg/g) [40], goethite-P (AAm) composite (1.22 mg/g) [13], iron-oxide-based adsorbents μGFH (22.4 mg/g) and μTMF (15.4 mg/g) [41], Fe-sericite composite beads (5.78 mg/g) [42], and red mud-modified biochar (5.923 mg/g) [43].
The adsorption capacity of PC-Fe/C-B was, thus, found to be 21.0 mg/g comparable to, but lower than, the following adsorbents: aluminium-modified crop-straw-derived biochars (34.71–52.02 mg/g) [44], zeo-NaY-S (37.59 mg/g) [45], magnetic nanocomposite (250 mg/g) [46], mesoporous-sulphated zirconia (99.23 mg/g) [47], polystyrene-Fe3O4 (139.3 mg/g) [48], and potassium ferrate (162.02 mg/g) [12]. In general, the smaller the particle size is, the larger the specific surface area and the higher the adsorption capacity. However, most adsorbents with large adsorption capacities are nanoparticles, and these are both unsuitable for fixed-bed-column adsorption and difficult to recycle. Moreover, nanoparticles easily agglomerate. PC-Fe/C-B can be easily ground into particles of a demanding size that are more suitable than nanoparticles for use in fixed-bed columns, for adsorptive As(V) removal [49]. Furthermore, most of the reported adsorption capacities were based on the results of static adsorption experiments. Compared with static adsorption [20], the diffusion condition in the flow state is superior in dynamic adsorption, and this enables full contact between the arsenate and adsorbent in the solution. Therefore, a dynamic experiment was more useful than a batch experiment, for evaluating the adsorption performance of the adsorbent. Similar to a previous study [50], the higher adsorption capacity of PC-Fe/C-B was attributed to the longer contact time between the adsorption material and the solution as well as better water-conservancy-diffusion conditions.

3.4. CD-MUSIC Modeling

3.4.1. Simulation Parameters of CD-MUSIC Model with PHREEQC

The CD-MUSIC model is an SCM developed from the classical TLM model, for the complexation of heavy-metal ions in hydrated oxides. Iron oxide combines metals and protons at both strong and weak sites, and the charges, potentials, and types of adsorbates are distributed in a double-layer plane, of more than two layers; therefore, its mass transfer region extends from the surface of the adsorbent to the free solution (Figure 11). There are three major differences between the CD-MUSIC and TLM models: the representation of surface acidity, the placement of ions and charges in the electrostatic plane, and the representation of reactive surface adsorption sites [51]. In the CD-MUSIC model, protonated surface groups are located in the 0-plane, charges of specific adsorbed ions (such as arsenate) are distributed between the 0-plane and 1-plane, and supporting electrolyte ions are assumed to be single-point charges located in the 1-plane. The diffuse layer starts from the 2-plane and extends to the bulk solution. C1 and C2 are the key parameters in the model, and these are 0.85 F/m2 and 0.75 F/m2, respectively [18].
CD-MUSIC is by far the most flexible model. However, it has one limitation, which is that many parameters need to be determined, and they are not equally constrained [52]. However, the CD-MUSIC model accurately characterised the surface-charge characteristics of PC-Fe/C-B, and the surface-charge characteristic parameters obtained by fitting are listed in Table 1 and Table 2. These are both the PC-Fe/C-B properties and the basic parameters for the adsorption-complex-reaction simulation.
According to a study by Hiemstra et al. (1996) [21], octahedrally coordinated iron-oxide surfaces mainly contain three types of surface sites: singly (≡FeO), doubly (≡Fe2O), and triply (≡Fe3O) coordinated sites, as shown in Figure 12. There are three types of Fe-O surface hydroxyl sites for iron oxides, namely ≡FeOH2+, ≡FeOH, and ≡FeO, but the unit-bound oxygen ≡FeOH2+ in the outer layer is, mainly, complexed with adsorbate-functional groups, such as As-O [21,53,54]. Combination forms include monodentate, bidentate, and tridentate combinations. The surface protonation reaction is defined by Equations (1) and (2), and the model parameters and ion-pair formation reactions for the singly coordinated groups are listed in Table 2.
FeOH 2 + = FeOH + H +
K 1 = [ FeOH ] [ H + ] / [ FeOH 2 + ]
It is usually considered that bidentate-surface complexes are inferior to monodentate complexes, for systems with a low As(V) coating coverage. However, Hayes et al. (1988) [55] reported simulation results for monodentate and bidentate surface species that were almost identical, and, thus, simpler surface species (monodentate) were used in this study to model As(V) adsorption [33,34].
The existing aqueous phase As(V) forms can be obtained from the following dissociation equations, and there are four possible As(V) surface reactions, in the formation of the above four aqueous As(V) species.
H + + AsO 4 3     HAsO 4 2
2 H + + AsO 4 3     H 2 AsO 4
3 H + + AsO 4 3     H 3 AsO 4 0
Of the four possible aqueous As(V) species, only two (H2AsO4 and HAsO42−) exist at typical pH values for drinking-water treatment (pH = 4–10). Therefore, in the SCF model simulation, it is necessary to consider the surface reactions given in Equations (9) and (11) (please see the equations provided below). When modelling the As(V)-FeAC system with the SCF model [33,34], the adsorption balance described using Equations (9) and (11) is referred to as the 2-RXN model. The 3-RXN model considers the effects of three surface reactions, when describing the adsorption balance, and Equation (7) (rather than Equation (13)) was selected as the third reaction (again, please see equations listed below). The removal rate of As(V) in a low-pH environment is generally the highest [56]. Although as As(V) is exclusively adsorbed on the oxide surface, by a ligand-exchange mechanism, and exists as a complex on the inner-sphere surface [33,34], four As(V) surface reactions are possible in the aqueous phase.
The column adsorption results were simulated by a transport–adsorptive reaction, using the PHREEQC program, and the surface-site density of PC-Fe/C-B was 2.50 sites/nm2. Xie et al. [57] studied the adsorption of Cr(VI) on goethite, using the CD-MUSIC model, and estimated the site density to be 2.7 sites/nm2, which is close to our estimation.
XOH 0 + H 3 AsO 4     XH 2 AsO 4 0 + H 2 O
K As 1 = XH 2 AsO 4 0 XOH   0 ] { H b + } AsO 4 3 exp 3 F Ψ RT
K As 2 = XHAsO 4 XOH   0 ] { H   b + } AsO 4 3 exp 2 F Ψ RT
XOH 0 + HAsO 4 2     XAsO 4 + H 2 O
K As 3 = XAsO 4 2 XOH   0 H b + AsO 4 3 exp F Ψ RT
XOH   0 + AsO 4 3     XOHAsO 4 3
K As 4 = XOHAsO 4 3 XOH   0 ] [ AsO 4 3
As mentioned above, the possible surface-complexation forms of PC-Fe/C-B-adsorbed As(V) included monodentate mononuclear, bidentate binuclear, and tridentate trinuclear. The adsorption reactions for As(V) in the CD-MUSIC-model simulation involved the following reactions: a surface-protonation reaction of the adsorption material ≡FeOH2+, a combination reaction for two electrolyte-ion pairs on the site of ≡FeOH2+, an arsenic acid three-stage dissociation reaction, and an adsorption reaction of AsO43- on the surface of PC-Fe/C-B. The basic parameters in Table 1 and Table 2 as well as the adsorption data under different conditions were entered into the software, to simulate the CD-MUSIC model. The f value (CD factor) of each surface morphology was manually adjusted, and the macro adsorption data were used as the constraint conditions of the model, by trial calculation, until appropriate results were obtained, as shown in Figure 13. The possible complex forms of AsO43− adsorbed by PC-Fe/C-B include the doubly protonated monodentate-surface complex ≡FeOAsO(OH)2, protonated monodentate-surface complex ≡FeOAsO2OH, non-protonated monodentate-surface complex ≡FeOAsO3, doubly protonated bidentate-surface complex ≡Fe2O2As(OH)2, protonated bidentate-surface complex ≡Fe2O2AsOOH, non-protonated monodentate-surface complex ≡Fe2O2AsO2, and non-protonated tridentate-surface complex ≡Fe3O3AsO.

3.4.2. Simulation Calculation of Solid-Surface-Complexing State

The adsorption behaviour of As(V) on PC-Fe/C-B at pH values of 1–13 was simulated using the CD-MUSIC model, and the changing trends of various forms under different pH conditions were predicted, as shown in Figure 13. The PHREEQC input files are detailed in the electronic Supplementary Material.
The simulation results in Figure 13 show that the possible complex forms of AsO43− adsorbed by PC-Fe/C-B were protonated monodentate-surface complex ≡FeOAsO2OH (complex constant of 26.62), protonated bidentate-surface complex ≡Fe2O2AsOOH (complex constant of 32.69) and non-protonated monodentate-surface complex ≡Fe2O2AsO2 (complex constant of 29.29). In a wide range of pH, the dominant form was protonated monodentate-surface complex ≡FeOAsO2OH, with the highest concentration at pH 4–7. At pH 9–12, the non-protonated monodentate-surface complex ≡Fe2O2AsO2 began to predominate. At pH > 10, the main form of the surface oxide was ≡FeO, which explained the low adsorption capacity of the adsorbent, under high-pH conditions.

3.4.3. Simulation of Reactive Transportation in Dynamic-Adsorption Process Using PHREEQC

As previously described, As adsorption has been extensively studied, using the CD-MUSIC model. Recently, Cui (2015) [58] used granular TiO2 to remove arsenic from groundwater in geogenic areas, and the As breakthrough curve was successfully calculated, using PHREEQC, incorporating the CD-MUSIC model and the 1D reactive transport block. In this study, the CD-MUSIC model was used to fit the experimental data of fixed-bed-column adsorption, and the fitting data were compared with those of the aforementioned classical dynamic model (see Figure 14). As shown in Figure 14, reactive transportation was simulated and calculated using the PHREEQC-coupled CD-MUSIC-surface-complexation model, of As(V) adsorbed by PC-Fe/C-B with the transport model of As(V) in aqueous solution. The breakthrough curve in the column-dynamic-adsorption process fitted well with the predicted data. In addition, the correlation coefficient (R2) was higher than that of the previous Thomas, Yoon–Nelson, Adams–Bohart, Clark, and Wolborska models. This satisfactory modelling result proves the effective application of CD-MUSIC, for predicting the column-breakthrough curves of As(V) adsorption in arsenic wastewater.

4. Conclusions

In this study, PC-Fe/C-B material was prepared, and the hierarchical porous microstructure of bamboo bio-templates was found to be accurately retained. The PC-Fe/C-B adsorbent was found to be effective for the removal of As(V) from wastewater, under the following optimal experimental conditions: influent flow 5.136 mL/min, pH 3, As(V) concentration 20 mg/L, adsorbent particle size 0.149 mm, adsorption temperature 35 °C, PC-Fe/C-B dose 0.5 g, and breakthrough time 50 min (184 BV), with qe,exp of 21.0 mg/g. The adsorption capacity of PC-Fe/C-B was higher than that of similar adsorbents reported in the literature.
The surface characteristics of PC-Fe/C-B were simulated and analysed, using the CD-MUSIC model with the PHREEQC program. The surface-site density of PC-Fe/C-B was estimated as 2.5 sites/nm2, and the surface complexes ≡FeOAsO2OH, ≡Fe2O2AsOOH, ≡Fe2O2AsO2, and ≡FeOAsO2OH were dominant over a wide pH range.
These results showed that the CD-MUSIC model can be effectively coupled with the transport model, using the PHREEQC program, to simulate the reactive transportation of As(V) in the fixed-bed column and to predict the breakthrough curve for column adsorption. The fitting result obtained was superior to that of classical models, such as the Thomas model and the Yoon–Nelson model.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/w14121848/s1. Table S1: Parameters of the Thomas model and the equilibrium As(V) uptake (qe) and the total removal percentage (Y) for As(V) adsorption onto PC-Fe/C-B under different conditions; Table S2: Parameters of the Yoon-–Nelson model, using linear-regression analysis for As(V) adsorption onto PC-Fe/C-B, under different conditions; Table S3: Parameters of the Clark model, using linear-regression analysis for As(V) adsorption onto PC-Fe/C-B, under different conditions; Table S4: Parameters of the Adams–Bohart and Wolborska models, using linear regression analysis for As(V) adsorption onto PC-Fe/C-B, under different conditions; Table S5: Comparison of the adsorption capacity of the PC-Fe/C-B and various adsorbents for As(V) removal from water; Table S6: Input files for the As(V) aqueous-phase speciation in the simulation, using PHREEQC; Table S7: PHREEQC input files for the surface acid–base behaviour of PC-Fe/C-B, in aqueous As(V) solution; Table S8: An input file example of the CD-MUSIC model for surface complexation in the simulation, using PHREEQC; Table S9: An input file example for the influent TRANSPORT in the simulation, using PHREEQC.

Author Contributions

Conceptualization, Y.Z. (Yinian Zhu) and Y.L.; methodology, Y.P. and L.Z.; software, Y.Z. (Yinian Zhu) and Y.L.; validation, Y.P.; formal analysis, J.Z.; investigation, Y.Z. (Yao Zhao); resources, X.Z. and Y.L.; data curation, S.T.; writing–original draft preparation, Y.P.; writ-ing–review and editing, Y.L.; visualization, L.Z.; supervision, S.T.; project administration, X.Z.; funding acquisition, Y.Z. (Yinian Zhu) and X.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Science Foundation of China (No. 52070050) and the Natural Science Foundation of Guangxi (No. 2020GXNSFAA159017).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data used in this study are available on request from the corresponding author.

Acknowledgments

The authors thank Guilin University of Technology for providing the materials, equipment, and laboratory facilities required to successfully conclude this research. All authors would like to sincerely thank the anonymous reviewers for the improvement of this paper, through their constructive and insightful comments.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Samuel, M.S.; Selvarajan, E.; Sarswat, A.; Muthukumar, H.; Jacob, J.M.; Mukesh, M.; Pugazhendhi, A. Nanomaterials as adsorbents for As (III) and As (V) removal from water: A review. J. Hazard. Mater. 2022, 424, 127572. [Google Scholar] [CrossRef] [PubMed]
  2. Purohit, S.; Chini, M.K.; Chakraborty, T.; Yadav, K.L.; Satapathi, S. Rapid removal of arsenic from water using metal oxide doped recyclable cross-linked chitosan cryogel. SN Appl. Sci. 2020, 2, 768. [Google Scholar] [CrossRef] [Green Version]
  3. Carneiro, M.A.; Pintor, A.; Boaventura, R.A.; Botelho, C. Current trends of arsenic adsorption in continuous mode: Literature review and future perspectives. Sustainability 2021, 13, 1186. [Google Scholar] [CrossRef]
  4. Liao, T.; Xi, Y.; Zhang, L.; Li, J.; Cui, K. Removal of toxic arsenic (As (Ⅲ)) from industrial wastewater by ultrasonic enhanced zero-valent lead combined with CuSO4. J. Hazard. Mater. 2021, 408, 124464. [Google Scholar] [CrossRef]
  5. Kumar, A.; Bhattacharya, T.; Shaikh, W.A.; Roy, A.; Mukherjee, S.; Kumar, M. Performance evaluation of crop residue and kitchen waste-derived biochar for eco-efficient removal of arsenic from soils of the Indo-Gangetic plain: A step towards sustainable pollution management. Environ. Res. 2021, 200, 111758. [Google Scholar] [CrossRef]
  6. Yang, X.; Li, J.; Liang, T.; Yan, X.; Zhong, L.; Shao, J.; El-Naggar, A.; Guan, C.; Liu, J.; Zhou, Y. A combined management scheme to simultaneously mitigate As and Cd concentrations in rice cultivated in contaminated paddy soil. J. Hazard. Mater. 2021, 416, 125837. [Google Scholar] [CrossRef]
  7. Lee, Y.; Ren, Y.; Cui, M.; Zhou, Y.; Kwon, O.; Ko, J.; Khim, J. Arsenic adsorption study in acid mine drainage using fixed bed column by novel beaded adsorbent. Chemosphere 2022, 291, 132894. [Google Scholar] [CrossRef]
  8. Glass, S.; Mantel, T.; Appold, M.; Sen, S.; Usman, M.; Ernst, M.; Filiz, V. Amine-terminated PAN membranes as anion-adsorber materials. Chem. Ing. Tech. 2021, 93, 1396–1400. [Google Scholar] [CrossRef]
  9. Asere, T.G.; Stevens, C.V.; Du Laing, G. Use of (modified) natural adsorbents for arsenic remediation: A review. Sci. Total Environ. 2019, 676, 706–720. [Google Scholar] [CrossRef]
  10. Gu, Y.; Xie, D.; Wang, Y.; Qin, W.; Zhang, H.; Wang, G.; Zhang, Y.; Zhao, H. Facile fabrication of composition-tunable Fe/Mg bimetal-organic frameworks for exceptional arsenate removal. Chem. Eng. J. 2019, 357, 579–588. [Google Scholar] [CrossRef]
  11. Ali, S.; Rizwan, M.; Shakoor, M.B.; Jilani, A.; Anjum, R. High sorption efficiency for As (III) and As (V) from aqueous solutions using novel almond shell biochar. Chemosphere 2020, 243, 125330. [Google Scholar] [CrossRef]
  12. Lan, B.; Wang, Y.; Wang, X.; Zhou, X.; Kang, Y.; Li, L. Aqueous arsenic (As) and antimony (Sb) removal by potassium ferrate. Chem. Eng. J. 2016, 292, 389–397. [Google Scholar] [CrossRef]
  13. Ramirez-Muñiz, K.; Perez-Rodriguez, F.; Rangel-Mendez, R. Adsorption of arsenic onto an environmental friendly goethite-polyacrylamide composite. J. Mol. Liq. 2018, 264, 253–260. [Google Scholar] [CrossRef]
  14. Navarathna, C.M.; Karunanayake, A.G.; Gunatilake, S.R.; Pittman, C.U., Jr.; Perez, F.; Mohan, D.; Mlsna, T. Removal of Arsenic (III) from water using magnetite precipitated onto Douglas fir biochar. J. Environ. Manag. 2019, 250, 109429. [Google Scholar] [CrossRef]
  15. Chen, M.L.; Sun, Y.; Huo, C.B.; Liu, C.; Wang, J.H. Akaganeite decorated graphene oxide composite for arsenic adsorption/removal and its proconcentration at ultra-trace level. Chemosphere 2015, 130, 52–58. [Google Scholar] [CrossRef]
  16. Usman, M.; Zarebanadkouki, M.; Waseem, M.; Katsoyiannis, I.A.; Ernst, M. Mathematical modeling of arsenic (V) adsorption onto iron oxyhydroxides in an adsorption-submerged membrane hybrid system. J. Hazard. Mater. 2020, 400, 123221. [Google Scholar] [CrossRef]
  17. Wang, Z.; Ma, T.; Zhu, Y.; Abass, O.K.; Liu, L.; Su, C.; Shan, H. Application of siderite tailings in water-supply well for As removal: Experiments and field tests. Int. Biodeterior. Biodegrad. 2018, 128, 85–93. [Google Scholar] [CrossRef]
  18. Stachowicz, M.; Hiemstra, T.; Riemsdijk, W. Surface speciation of As(III) and As(V) in relation to charge distribution. J. Colloid Interface Sci. 2006, 302, 62–75. [Google Scholar] [CrossRef]
  19. Huang, W.; Xu, H.; Liu, X.; Wang, L.; Li, S.; Ji, L.; Qu, Z.; Yan, N. Surface protection method for the magnetic core using covalent organic framework shells and its application in As (III) depth removal from acid wastewater. J. Environ. Sci. 2022, 115, 1–9. [Google Scholar] [CrossRef]
  20. Zhu, Y.; Zhu, Z.; Chen, Y.; Yang, F.; Hui, Q. Kinetics and thermodynamics of sorption for As(V) on the porous biomorph-genetic composite of α-Fe2O3/Fe3O4/C with eucalyptus wood hierarchical microstructure. Water Air Soil Pollut. 2013, 224, 1589. [Google Scholar] [CrossRef]
  21. Hiemstra, T.; Van Riemsdijk, W.H. A surface structural approach to ion adsorption: The charge distribution (CD) model. J. Colloid Interface Sci. 1996, 179, 488–508. [Google Scholar] [CrossRef]
  22. Deng, Y.; Li, Y.; Li, X.; Wang, L.; Ma, J.; Chen, Y.; Weng, L. Development and application of chemical speciation models for heavy metals in environmental soil samples. J. Agro-Environ. Sci. 2018, 37, 1350–1361. [Google Scholar]
  23. Salazar-Camacho, C.; Villalobos, M. Goethite surface reactivity: III. Unifying arsenate adsorption behavior through a variable crystal face-site density model. Geochim. Cosmochim. Acta 2010, 74, 2257–2280. [Google Scholar] [CrossRef]
  24. Gustafsson, J.P. Modelling competitive anion adsorption on oxide minerals and an allophane-containing soil. Eur. J. Soil Sci. 2001, 52, 639–653. [Google Scholar] [CrossRef]
  25. Liang, Y.; Xu, J.; Koopal, L.K.; Wang, M.; Xiong, J.; Hou, J.; Tan, W. Facet-dependent surface charge and Pb2+ adsorption characteristics of hematite nanoparticles: CD-MUSIC-eSGC modeling. Environ. Res. 2021, 196, 110383. [Google Scholar] [CrossRef]
  26. Parkhurst, D.L.; Appelo, C. Description of Input and Examples for PHREEQC Version 3: A Computer Program for Speciation, Batch-Reaction, One-Dimensional Transport, and Inverse Geochemical Calculations; US Geological Survey: Reston, VA, USA, 2013; pp. 2328–7055. [Google Scholar]
  27. Thomas, H.C. Heterogeneous ion exchange in a flowing system. J. Am. Chem. Soc. 1944, 66, 1664–1666. [Google Scholar] [CrossRef]
  28. Yoon, Y.H.; Nelson, J.H. Application of gas adsorption kinetics I. A theoretical model for respirator cartridge service life. Am. Ind. Hyg. Assoc. J. 1984, 45, 509–516. [Google Scholar] [CrossRef]
  29. Yoon, Y.H.; Nelson, J.H. Application of gas adsorption kinetics—II. A theoretical model for respirator cartridge service life and its practical applications. Am. Ind. Hyg. Assoc. J. 1984, 45, 517–524. [Google Scholar] [CrossRef]
  30. Bohart, G.; Adams, E. Some aspects of the behavior of charcoal with respect to chlorine. J. Am. Chem. Soc. 1920, 42, 523–544. [Google Scholar] [CrossRef] [Green Version]
  31. Clark, R.M. Evaluating the cost and performance of field-scale granular activated carbon systems. Environ. Sci. Technol. 1987, 21, 573–580. [Google Scholar] [CrossRef]
  32. Wolborska, A.; Pustelnik, P. A simplified method for determination of the break-through time of an adsorbent layer. Water Res. 1996, 30, 2643–2650. [Google Scholar] [CrossRef]
  33. Vaughan, R.L., Jr.; Reed, B.E. Modeling As (V) removal by a iron oxide impregnated activated carbon using the surface complexation approach. Water Res. 2005, 39, 1005–1014. [Google Scholar] [CrossRef]
  34. Vaughan, R.L., Jr.; Reed, B.E.; Smith, E.H. Modeling As (V) removal in iron oxide impregnated activated carbon columns. J. Environ. Eng. 2007, 133, 121–124. [Google Scholar] [CrossRef]
  35. Vithanage, M.; Rajapaksha, A.U.; Dou, X.; Bolan, N.S.; Yang, J.E.; Ok, Y.S. Surface complexation modeling and spectroscopic evidence of antimony adsorption on iron-oxide-rich red earth soils. J. Colloid Interface Sci. 2013, 406, 217–224. [Google Scholar] [CrossRef]
  36. Moulder, J.; Stickle, W.; Sobol, P.; Bomben, K. Handbook of X-ray Photoelectron Spectroscopy; Chastain, J., Ed.; Perkin-Elmer Corporation: Waltham, MA, USA, 1992. [Google Scholar]
  37. Costa, E.; Guilherme, L.; Lopes, G.; Lima, J.; Curi, N. Competitive sorption of arsenate and phosphate on aluminum mining by-product. Water Air Soil Pollut. 2012, 223, 5433–5444. [Google Scholar] [CrossRef]
  38. Misaelides, P.; Nikashina, V.; Godelitsas, A.; Gembitskii, P.; Kats, E. Sorption of As(V)-anions from aqueous solutions by organo-modified natural zeolitic materials. J. Radioanal. Nucl. Chem. 1998, 227, 183–186. [Google Scholar] [CrossRef]
  39. Luo, L.; Zhang, S.; Shan, X.; Zhu, Y. Effects of oxalate and humic acid on arsenate sorption by and desorption from a Chinese red soil. Water Air Soil Pollut. 2006, 176, 269–283. [Google Scholar] [CrossRef] [Green Version]
  40. Vieira, B.R.; Pintor, A.M.; Boaventura, R.A.; Botelho, C.M.; Santos, S.C. Arsenic removal from water using iron-coated seaweeds. J. Environ. Manag. 2017, 192, 224–233. [Google Scholar] [CrossRef]
  41. Usman, M.; Belkasmi, A.I.; Kastoyiannis, I.A.; Ernst, M. Pre-deposited dynamic membrane adsorber formed of microscale conventional iron oxide-based adsorbents to remove arsenic from water: Application study and mathematical modeling. J. Chem. Technol. Biotechnol. 2021, 96, 1504–1514. [Google Scholar] [CrossRef]
  42. Lee, C.; Jung, J.; Pawar, R.R.; Kim, M.; Lee, S.M. Arsenate and phosphate removal from water using Fe-sericite composite beads in batch and fixed-bed systems. J. Ind. Eng. Chem. 2017, 47, 375–383. [Google Scholar] [CrossRef]
  43. Wu, C.; Huang, L.; Xue, S.; Huang, Y.; Hartley, W.; Cui, M.; Wong, M. Arsenic sorption by red mud-modified biochar produced from rice straw. Environ. Sci. Pollut. Res. 2017, 24, 18168–18178. [Google Scholar] [CrossRef]
  44. Qian, W.; Zhao, A.; Xu, R. Sorption of As (V) by aluminum-modified crop straw-derived biochars. Water Air Soil Pollut. 2013, 224, 1610. [Google Scholar] [CrossRef]
  45. Yusof, A.M.; Malek, N.A.N.N. Removal of Cr (VI) and As (V) from aqueous solutions by HDTMA-modified zeolite Y. J. Hazard. Mater. 2009, 162, 1019–1024. [Google Scholar] [CrossRef]
  46. Alijani, H.; Shariatinia, Z. Effective aqueous arsenic removal using zero valent iron doped MWCNT synthesized by in situ CVD method using natural α-Fe2O3 as a precursor. Chemosphere 2017, 171, 502–511. [Google Scholar] [CrossRef]
  47. Han, L.; Sun, H.; Ro, K.S.; Sun, K.; Libra, J.A.; Xing, B. Removal of antimony (III) and cadmium (II) from aqueous solution using animal manure-derived hydrochars and pyrochars. Bioresour. Technol. 2017, 234, 77–85. [Google Scholar] [CrossRef] [Green Version]
  48. Jiang, W.; Chen, X.; Niu, Y.; Pan, B. Spherical polystyrene-supported nano-Fe3O4 of high capacity and low-field separation for arsenate removal from water. J. Hazard. Mater. 2012, 243, 319–325. [Google Scholar] [CrossRef]
  49. Luo, T.; Cui, J.; Hu, S.; Huang, Y.; Jing, C. Arsenic removal and recovery from copper smelting wastewater using TiO2. Environ. Sci. Technol. 2010, 44, 9094–9098. [Google Scholar] [CrossRef]
  50. Zhang, L.; Zeng, Y.; Cheng, Z. Removal of heavy metal ions using chitosan and modified chitosan: A review. J. Mol. Liq. 2016, 214, 175–191. [Google Scholar] [CrossRef]
  51. Mangold, J.E.; Park, C.M.; Liljestrand, H.M.; Katz, L.E. Surface complexation modeling of Hg (II) adsorption at the goethite/water interface using the Charge Distribution Multi-Site Complexation (CD-MUSIC) model. J. Colloid Interface Sci. 2014, 418, 147–161. [Google Scholar] [CrossRef]
  52. Sø, H.U.; Postma, D.; Jakobsen, R.; Larsen, F. Competitive adsorption of arsenate and phosphate onto calcite; experimental results and modeling with CCM and CD-MUSIC. Geochim. Cosmochim. Acta 2012, 93, 1–13. [Google Scholar] [CrossRef]
  53. Antelo, J.; Avena, M.; Fiol, S.; López, R.; Arce, F. Effects of pH and ionic strength on the adsorption of phosphate and arsenate at the goethite–water interface. J. Colloid Interface Sci. 2005, 285, 476–486. [Google Scholar] [CrossRef] [PubMed]
  54. Antelo, J.; Arce, F.; Fiol, S. Arsenate and phosphate adsorption on ferrihydrite nanoparticles. Synergetic interaction with calcium ions. Chem. Geol. 2015, 410, 53–62. [Google Scholar] [CrossRef]
  55. Hayes, K.F.; Papelis, C.; Leckie, J.O. Modeling ionic strength effects on anion adsorption at hydrous oxide/solution interfaces. J. Colloid Interface Sci. 1988, 125, 717–726. [Google Scholar] [CrossRef]
  56. Dzombak, D.A.; Morel, F.M. Surface Complexation Modeling: Hydrous Ferric Oxide; John Wiley & Sons: Hoboken, NJ, USA, 1990. [Google Scholar]
  57. Xie, J.; Gu, X.; Tong, F.; Zhao, Y.; Tan, Y. Surface complexation modeling of Cr(VI) adsorption at the goethite–water interface. J. Colloid Interface Sci. 2015, 455, 55–62. [Google Scholar] [CrossRef]
  58. Cui, J.; Du, J.; Yu, S.; Jing, C.; Chan, T. Groundwater arsenic removal using granular TiO2: Integrated laboratory and field study. Environ. Sci. Pollut. Res. 2015, 22, 8224–8234. [Google Scholar] [CrossRef]
Figure 1. SEM images of PC-Fe/C-B, after As(V) adsorption: (a) transverse section. (b) longitudinal section and (c) EDS spectrum of PC-Fe/C-B, after As(V) adsorption.
Figure 1. SEM images of PC-Fe/C-B, after As(V) adsorption: (a) transverse section. (b) longitudinal section and (c) EDS spectrum of PC-Fe/C-B, after As(V) adsorption.
Water 14 01848 g001
Figure 2. XRD patterns of PC-Fe/C-B, before and after adsorption.
Figure 2. XRD patterns of PC-Fe/C-B, before and after adsorption.
Water 14 01848 g002
Figure 3. FT-IR patterns of PC-Fe/C-B, before and after As(V) adsorption.
Figure 3. FT-IR patterns of PC-Fe/C-B, before and after As(V) adsorption.
Water 14 01848 g003
Figure 4. (a) XPS survey spectra and high-resolution. (b) C 1s. (c) O 1s. (d) Fe 2p and (e) As 3d spectra of PC-Fe/C-B, before and after As(V) adsorption.
Figure 4. (a) XPS survey spectra and high-resolution. (b) C 1s. (c) O 1s. (d) Fe 2p and (e) As 3d spectra of PC-Fe/C-B, before and after As(V) adsorption.
Water 14 01848 g004
Figure 5. Effect of influent concentration on breakthrough curve of PC-Fe/C-B adsorption As(V).
Figure 5. Effect of influent concentration on breakthrough curve of PC-Fe/C-B adsorption As(V).
Water 14 01848 g005
Figure 6. Effect of influent flow rate on breakthrough curve of PC-Fe/C-B adsorption As(V).
Figure 6. Effect of influent flow rate on breakthrough curve of PC-Fe/C-B adsorption As(V).
Water 14 01848 g006
Figure 7. Effect of influent pH on breakthrough curve of PC-Fe/C-B adsorption As(V).
Figure 7. Effect of influent pH on breakthrough curve of PC-Fe/C-B adsorption As(V).
Water 14 01848 g007
Figure 8. Effect of adsorbent dose on breakthrough curve of PC-Fe/C-B adsorption As(V).
Figure 8. Effect of adsorbent dose on breakthrough curve of PC-Fe/C-B adsorption As(V).
Water 14 01848 g008
Figure 9. Effect of adsorbent grain size on breakthrough curve of PC-Fe/C-B adsorption As(V).
Figure 9. Effect of adsorbent grain size on breakthrough curve of PC-Fe/C-B adsorption As(V).
Water 14 01848 g009
Figure 10. Effect of temperature on breakthrough curve of PC-Fe/C-B adsorption As(V).
Figure 10. Effect of temperature on breakthrough curve of PC-Fe/C-B adsorption As(V).
Water 14 01848 g010
Figure 11. Schematic representation of the parameters of the CD-MUSIC adsorption model.
Figure 11. Schematic representation of the parameters of the CD-MUSIC adsorption model.
Water 14 01848 g011
Figure 12. Charge on singly, doubly, and triply coordinated-surface hydroxyl -functional groups, in octahedrally coordinated iron oxide.
Figure 12. Charge on singly, doubly, and triply coordinated-surface hydroxyl -functional groups, in octahedrally coordinated iron oxide.
Water 14 01848 g012
Figure 13. Speciation distribution of surface complexation on PC-Fe/C-B, after As(V) adsorption (initial As = 20 mg/L).
Figure 13. Speciation distribution of surface complexation on PC-Fe/C-B, after As(V) adsorption (initial As = 20 mg/L).
Water 14 01848 g013
Figure 14. Experimental data and coupled CD-MUSIC–transport simulation, using PHREEQC, for PC-Fe/C-B adsorption As(V) (initial concentration of As(V) = 20 mg/L, pH = 3, Q = 5.136 mL/min, dose m = 0.5 g, particle size < 0.149 mm, T = 35 °C).
Figure 14. Experimental data and coupled CD-MUSIC–transport simulation, using PHREEQC, for PC-Fe/C-B adsorption As(V) (initial concentration of As(V) = 20 mg/L, pH = 3, Q = 5.136 mL/min, dose m = 0.5 g, particle size < 0.149 mm, T = 35 °C).
Water 14 01848 g014
Table 1. Speciation of As(V) in aqueous phase.
Table 1. Speciation of As(V) in aqueous phase.
As(V) Aqueous-Phase-Dissociation ReactionlogK
H3AsO4 = AsO43− + 3H+−20.7
H+ + AsO43− = HAsO42−11.5
2 H+ + AsO43− = H2AsO418.46
Table 2. Surface-complexation-equilibrium reactions of As(V) adsorption on PC-Fe/C-B, and the parameters used in CD-MUSIC model simulation.
Table 2. Surface-complexation-equilibrium reactions of As(V) adsorption on PC-Fe/C-B, and the parameters used in CD-MUSIC model simulation.
Surface Reaction of Adsorption of AsO43− by PC-Fe/C-BlogK
≡FeOH = ≡FeO + H+−8.93
≡FeOH + H+ = ≡FeOH2+7.29
2 ≡FeOH + 2 H+ + AsO43− = ≡Fe2O2AsO2 + 2 H2O29.29
2 ≡FeOH + 3 H+ + AsO43− = ≡Fe2O2AsOOH + 2 H2O32.69
≡FeOH + 2 H+ + AsO43− = ≡FeOAsO2OH + H2O26.62
The surface site density of PC-Fe/C-B/(sites/nm2): 2.5
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Peng, Y.; Li, Y.; Tang, S.; Zhang, L.; Zhang, J.; Zhao, Y.; Zhang, X.; Zhu, Y. Dynamic Adsorption of As(V) onto the Porous α-Fe2O3/Fe3O4/C Composite Prepared with Bamboo Bio-Template. Water 2022, 14, 1848. https://doi.org/10.3390/w14121848

AMA Style

Peng Y, Li Y, Tang S, Zhang L, Zhang J, Zhao Y, Zhang X, Zhu Y. Dynamic Adsorption of As(V) onto the Porous α-Fe2O3/Fe3O4/C Composite Prepared with Bamboo Bio-Template. Water. 2022; 14(12):1848. https://doi.org/10.3390/w14121848

Chicago/Turabian Style

Peng, Yuqing, Yanhong Li, Shen Tang, Lihao Zhang, Jing Zhang, Yao Zhao, Xuehong Zhang, and Yinian Zhu. 2022. "Dynamic Adsorption of As(V) onto the Porous α-Fe2O3/Fe3O4/C Composite Prepared with Bamboo Bio-Template" Water 14, no. 12: 1848. https://doi.org/10.3390/w14121848

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