5.2
Impact Factor
Generic selectors
Exact matches only
Search in title
Search in content
Post Type Selectors
Search in posts
Search in pages
Filter by Categories
Corrigendum
Current Issue
Editorial
Erratum
Full Length Article
Full lenth article
Letter to Editor
Original Article
Research article
Retraction notice
Review
Review Article
SPECIAL ISSUE: ENVIRONMENTAL CHEMISTRY
5.3
Impact Factor
Generic selectors
Exact matches only
Search in title
Search in content
Post Type Selectors
Search in posts
Search in pages
Filter by Categories
Corrigendum
Current Issue
Editorial
Erratum
Full Length Article
Full lenth article
Letter to Editor
Original Article
Research article
Retraction notice
Review
Review Article
SPECIAL ISSUE: ENVIRONMENTAL CHEMISTRY
View/Download PDF

Translate this page into:

Original article
12 (
8
); 4675-4693
doi:
10.1016/j.arabjc.2016.08.006

Arsenic removal by magnetite-loaded amino modified nano/microcellulose adsorbents: Effect of functionalization and media size

Faculty of Technology and Metallurgy, University of Belgrade, Karnegijeva 4, 11120 Belgrade, Serbia
Vinča Institute of Nuclear Sciences, University of Belgrade, 11000 Belgrade, Serbia
Military Academy, 33 General Pavle Jurišić – Šturm Street, Belgrade, Serbia
Innovation Center of the Faculty of Technology and Metallurgy, University of Belgrade, Karnegijeva 4, 11120 Belgrade, Serbia
Faculty of Forestry, University of Belgrade, Kneza Višeslava 1, 11030 Belgrade, Serbia
Faculty of Agriculture, University of Belgrade, Nemanjina 6, 11080 Belgrade-Zemun, Serbia
Institute of Technical Sciences of the SASA, Belgrade, Serbia

⁎Corresponding author. marinko@tmf.bg.ac.rs (Aleksandar Marinković)

Disclaimer:
This article was originally published by Elsevier and was migrated to Scientific Scholar after the change of Publisher.

Peer review under responsibility of King Saud University.

Abstract

Comparative adsorption study related to benefits of parent media size, i.e. microfibrillated cellulose (MC) versus nanocellulose (NC) support, for the preparation of magnetite (MG) based high performance adsorbent for arsenic removal was conducted. Precipitation of MG on amino terminal branched organic structure, L, either linked by maleic acid residue on NC surface (NC-MA/L) or linked by oxalyl bridge on MC surface (MC-O/L) produced NC-MA/L-MG and MC-O/L-MG adsorbents, respectively. Precipitation of nanosized MG on amino functionalized NC-MA/L and MC-O/L, performed according to optimized procedure, contributed to improved textural properties and adsorptive/kinetic performances of novel adsorbents. Adsorption capacity of arsenate, As(V), was in favor of NC-MA/L-MG (85.3 versus 18.5 mg g−1) while MC-O/L-MG exhibited faster kinetics (0.541 versus 0.189 g mg−1 min−1). Lower capacity of arsenite, As(III), removal, 68.3 mg g−1 for NC-MA/L-MG and 17.8 mg g−1 for MC-O/L-MG, were obtained. Calculated activation energies, 13.28 and 10.87 kJ mol−1 for NC-MA/L-MG and MC-O/L-MG with respect to As(V), respectively, suggest, in accordance with results of Weber-Morris fitting, that internal mass transfer controls adsorption process. Model free adsorption kinetics confirmed beneficial uses of MC-O/L-MG due to low activation energy dependence on the extent of adsorption.

Keywords

Sorption
Nanocellulose
Microcellulose
Arsenic
Magnetite
1

1 Introduction

Arsenic is classified as a Class A human carcinogen, toxic and mutagen and its contamination of drinking water has been a challenge of global magnitude (Ravenscroft et al., 2009). Due to that, exposure to arsenic can cause several skin problems such as pigmentation and hard skin patches, as well as bladder, kidney, lung, liver and prostate cancers (Awual et al., 2012). To limit the exposure to arsenic, the United State Environmental Protection Agency and the World Health Organization promulgated new water standards and decreased the maximum contaminant level (MCL) of arsenic to 10 μg L−1 (Ravenscroft et al., 2009). Subsequently, the strict regulation has initiated extensive development of technologies for effective arsenic removal.

Nowadays, developing of techniques for arsenic removal from water has received extensive attention by environmental scientists. Arsenic removal techniques include coagulation, oxidation, adsorption and reverse osmosis processes, and ion/ligand exchangers (Awual and Jyo, 2009; Awual et al., 2013, 2012, 2011, 2008). Among the possible arsenic removal techniques, adsorption is considered as simple, efficient and economic method which offers flexibility in design and generating high-quality treated effluent. Different adsorbents such as activated carbons, biological materials and mineral oxides were applied (Yousif et al., 2016). In the recent years, ferric (hydr)oxides have attached growing interest in arsenic removal due to their selectivity, chemical stability, low price, availability, and possibility to be produced as nanomaterials with large surface area. In most cases, however, the applicability of the ferric (hydr)oxides nanoparticles is shown to suffer from their poor mechanical strength and tendency to agglomerate. Application of support and production of hybrid materials has emerged as viable option in the preservation of beneficial properties of ferric (hydr)oxide nanomaterials (Hua et al., 2012).

The use of adsorbents containing natural polymers for heavy metals removal through adsorption takes particular attention (Abdel-Halim and Al-Deyab, 2011; Huang et al., 2013; Podder and Majumder, 2019; Sheng et al., 2014; Yousif et al., 2016; Zeng et al., 2015). From the ecological point of view, cellulose is biocompatible, the most widely available and renewable biopolymer available from natural sources (Lavoine et al., 2012). Considering that chemical transformation of hydroxyl groups plays an important role in enhancing the sorption capacity, cellulose proved to be an excellent candidate as support matrix amenable for surface modification (Lavoine et al., 2012; Šimkovic, 2013). Introduction of terminal amino and carboxylic groups provides effective means for ferric (hydr)oxides precipitation. Furthermore, the presence of amino moieties influences the synthesis/precipitation of ferric oxides and affects the particle size distribution and the colloidal stability in solution (Schwaminger et al., 2015). Functionalized nanosized (Anirudhan and Jalajamony, 2010; Anirudhan et al., 2012; Anirudhan et al., 2015; Donia et al., 2012; Singh et al., 2015; Sun et al., 2014; Yousif et al., 2016; Zhao et al., 2009), as well as microfibrillated cellulose (MC) was used for adsorption of As, F, Cr, Pb, Hg, Cu, Ag, U, and organic pollutants (Alila and Boufi, 2009; Aloulou et al., 2006; Hokkanen et al., 2016; Li et al., 2015; Maliyekkal et al., 2010; Tian et al., 2011; Yu et al., 2013). To our best knowledge, there are some reports about the cellulose/ferric (hydr)oxide hybrid materials used for arsenic removal (Zhao et al., 2009), whereas there is no report studying the influence of surface modification/type of functionalities on successfulness of ferric oxides precipitation, their properties and adsorption performances.

However, processing nanocellulose (NC) particles and application in a realistic adsorption systems is severely limited by a number of inherent shortcomings: purification, manipulation, and low solubility in organic solvents. Solubility improvement and deaggregation of cellulose particles could be overcome by designing of two general strategies: (1) the covalent functionalization of OH groups at NC surface with organic pendant group, and (2) the noncovalent functionalization through supramolecular interactions, which allows the formation of stable NC suspensions. More attractive and effective choice is surface covalent modification which could be conducted to introduce high affinity groups for binding/precipitation of the ion/oxides, e.g. ferric oxides applicable for arsenic removal. The valuable alternative is design of hierarchical organic structure with numerous terminal amino groups which provide support for efficient precipitation of ferric oxide, control of particle size and adsorbent integrity.

To achieve this goal, amino terminal branched organic structure, L, linked either by maleic acid residue on NC surface (NC-MA/L) or by oxalyl bridge on MC surface (MC-O/L) was used as support for magnetite (MG) precipitation. Design and synthesis of structure L, obtained by successive reactions of NC-MA and MC-O with ethylenediamine (EDA), ethylenediamine- tetraacetic acid (EDTA) ethyl ester and EDA in a final step, produced significant number of amino terminal groups. Nanosized MG was deposited/precipitated in an optimized/controlled manner on both NC-MA/L and MC-O/L surfaces producing NC-MA/L-MG and MC-O/L-MG adsorbents, respectively. The three main goals of presented study were devoted to elucidation of the benefits of parent media size, NC and MC, and the extent of the modification of these starting materials, i.e. introduction of the number of amino terminal functionalities, and controllable MG precipitation on adsorption performances of novel synthesized adsorbents. The specific objectives of the present study were related to comparison of the performances of both adsorbents in terms of the following: (1) textural and morphological properties; (2) adsorption efficiency; (3) kinetics/diffusional processes, and (4) adsorbent affinity to arsenic.

2

2 Experimental

2.1

2.1 Chemicals and materials

Cotton used for NC isolation was supplied from AD. Niva, Serbia (Turkish origin) and microfibrillated cellulose was supplied by Stora Enso Finland. All the chemicals used in this study were of analytical grade and used as received. The following chemicals were supplied from Sigma-Aldrich: sulfuric acid, glacial acetic acid (GAA), toluene, perchloric acid, maleic anhydride (MA), tetrahydrofurane (THF), ethylenediamine (EDA), ethylenediaminetetraacetic acid (EDTA), N,N-dimethylformamide (DMF), iron(II) sulfate heptahydrate (FeSO4·7H20), potassium nitrate (KNO3), potassium hydroxide (KOH), triethylamine (TEA), ethyl oxalyl chloride (EOC), and dicyclohexyl carbodiimide (DCC). The As(V) and As(III) stock solutions were prepared with deionized water (DW), resistivity 18 MΩ cm, using Na2HAsO4·7H2O (Sigma-Aldrich) and NaAsO2 p.a. (J.T. Baker, reagent grade), respectively. Sodium metasilicate nonahydrate (Na2SiO3·9H2O) (p.a.), potassium dichromate (K2Cr2O7) (p.a.) and sodium dihydrogen phosphate (NaH2PO4) (p.a.) reagent grade chemicals were purchased from Sigma-Aldrich.

2.2

2.2 Adsorbent preparation

The preparation of adsorbents was performed through five consecutive steps applying magnetic stirring (Heidolph) and ultrasonic treatment (Bandello electronic, Berlin, Germany, power 120 W, frequency 35 kHz) or combination of both methods. Such methodology was applied in order to enhance reactant diffusivity/reactivity due to limited availability of surface functional group being positioned at the internal adsorbent surface. By the use of ultrasound, the liquid circulation and related turbulence generated the formation and collapse of bubbles known as cavitation, thus inducing and improving mass and energy transfer. In this work, ultrasound assisted synthesis proved increased frequency of collision and enhanced reaction efficiency of some reaction steps. In that way, it offers an acceptable alternative for the synthesis of the adsorbents via a simple, catalyst-free reaction conditions which help in overcoming rate limiting steps in the course of constructing branched organic structure at NC and MC surface (Markovski et al., 2014a, 2014b).

2.3

2.3 Nanocellulose isolation

Following the procedure of sulfuric acid hydrolysis method (Lu and Hsieh, 2010) for nanocellulose isolation, the NC was produced from cotton microfiber (20 g) by treating with sulfuric acid (200 ml, 64 wt.%) at temperature < 40 °C for 1 h. Product was diluted with 800 ml of DW, neutralized to pH 6, and three cycles of centrifugation/DW water washing were applied. After dialyzing for 48 h, nanocellulose was filtered by using polytetrafluoroethylene (45 nm) filter membrane and denoted as NC.

2.4

2.4 NC modification

In order to achieve successful and controlled magnetite precipitation, as well as colloidal stability of cellulose based adsorbent, following amino modification steps were done:

  1. Modification with maleic anhydride (MA)NC (10 g) was dispersed in the mixture of glacial acetic acid (100 mL), toluene (130 mL) and perchloric acid (0.5 mL) by the use of ultrasound. Modification of NC was performed by dropwise addition of the solution of MA (5 g) in 45 mL of toluene for 15 min under sonication, and followed by maintaining reaction on a magnetic stirrer for 1 h at room temperature. Raw product was washed thoroughly with toluene and ethanol, and vacuum dried at 40 °C for 4 h. Obtained material was denoted NC-MA.

  2. NC-MA modification with EDA – Dispersion of NC-MA (10 g) in the 200 mL of tetrahydrofurane (THF) was added slowly into solution of EDA (10 mL) in THF (50 mL) in a three necked flask (500 mL) under ultrasonic treatment. After addition of 2 g of DCC dissolved in 20 mL of THF, the reaction was carried out by combination of ultrasonic/mixing treatment for 30 min (5 min/5 min sonication/mixing cycles), and continued with magnetic stirring at 20–25 °C for 3 h. Product was washed with THF and ethanol (three times), and then vacuum dried at 40 °C for 4 h. This material was denoted NC-MA/EDA.

  3. NC-MA/EDA modification with tetraethyl ester of EDTA – In the dispersion of 10 g of the NC-MA/EDA in 200 ml of absolute ethanol, ethanol solution of N, N′-1, 2-ethanediylbis[N-(2-ethoxy-2-oxoethyl)]-1,1′-glycine diethyl ester (tetraethyl ester of EDTA) (27 g) was added slowly. Tetraethyl ester of EDTA was prepared according to literature procedure (Jaeger et al., 2007). Reaction took place under ultrasonic treatment for 10 min, and followed with magnetic stirring at room temperature for 6 h. Product was washed with ethanol, and vacuum dried at 40 °C for 4 h. Obtained material was denoted NC-MA/EDA-EDTA.

  4. NC-MA/EDA-EDTA modification with EDANC-MA/EDA-EDTA (10 g) was subjected to ultrasound treatment in N,N-dimethylformamide (DMF) for 5 min, dropwise addition of EDA (10 mL) for 10 min was followed by heating/mixing at magnetic stirrer for 2 h at 50 °C. Obtained product was washed by using abundant quantity of DMF and DW, and dried at 60 °C for 12 h under vacuum. Synthesized material was denoted as NC-MA/EDA-EDTA-EDA (NC-MA/L).

2.5

2.5 Synthesis of NC-MA/EDA-EDTA-EDA-MG adsorbent (NC-MA/L-MG)

In order to achieve optimized method of NC-MA/L-MG synthesis and shorten experimental procedure the Response Surface Methodology (RSM) was applied for designing of the optimization of MG precipitation on NC-MA/L (Table S1; Supplementary data). The NC-MA/L (10 g) was dispersed in 100 mL FeSO4·7H2O solution of different concentrations: 0.013, 0.04, 0.1, 0.13, 0.16, 0.19, 0.23, 0.27, 0.32, 0.35, 0.39 and 0.49 mol L−1 performed ultrasonic treatment under inert atmosphere (N2) for 10 min. After sonication NC dispersion was transferred to oil bath, with constant gentle mixing, and temperature was adjusted at 90 °C using magnetic stirrer/heater. Afterward, 45 mL of oxygen-free solution containing different quantities of reactants: from 0.03 g to 1.1 g KNO3, and from 0.2 g to 7.6 g KOH, calculated according to literature method (Schwertmann and Cornell, 2000) with respect to starting FeSO4·7H2O, was added during 30 min providing continuous iron oxide precipitation. After precipitation, reaction was heated for additional 60 min at 90 °C, and left overnight. The adsorbent NC-MA/L-MG was isolated in the form of black precipitate, washed with DW until getting free from ions in filtrate and vacuum dried at 40 °C for 6 h. For the comparative purpose a sample of MG was prepared according to literature method (Schwertmann and Cornell, 2000).

2.6

2.6 Synthesis of MC-O/L-MG hybrid adsorbent

MC (10 g) was soaked in a mixture of dry THF (250 mL) and triethylamine (TEA) (30 mL) followed by dropwise addition of ethyl oxalyl chloride (EOC) (30 mL) in dry THF (50 mL). Reaction was sonicated for 1 h at room temperature. Isolated product ethyl oxalyl modified MC, MC-O, (10 g) was used as a media for linkage of EDA-EDTA-EDA structure in an analogous manner as it was performed for NC-MA/L. In that way, MC-O/L modified microfibrillated MC material was obtained. MC-O/L was subsequently modified with MG in the way analogous to preparation of NC-MA/L-MG adsorbent (Section 2.5). RSM experimental plan (Table S1; Supplementary data) was designed according to determined total basic sites at MC-O/L surface (0.60 mmol g−1; Section 3.1).

2.7

2.7 Solvent/nonsolvent (water/xylene) system applied for controllable impregnation of the NC-MA/L and MC-O/L with magnetite

The results of the optimization procedure were applied for the controllable impregnation of MG on NC-MA/L support in a solvent/nonsolvent system. NC (10 g) was soaked with xylene (non-solvent used to provide continual medium and wetting of the NC material; mixing of the medium was achieved by nitrogen/air bubbling in a upstream flow) in a perforated column with gas inlet/outlet valves, solvent exhaustion valve and system for vacuum drying of the NC at different modification steps. After establishment of continuous flow of nitrogen through two phase systems xylene/NC-MA/L and xylene/MC-O/L, slow addition (15 min) of the 40 mL (0.95 mol L−1) and 20 mL (0.85 mol L−1) of FeSO4·7H2O solution, respectively, started from dropping funnel situated at the top of column. Fine distribution of the FeSO4·7H2O solution was achieved by the glass frit assembled at the funnel. Small droplets of water iron solution were attracted by hydrophilic NC-MA/L and MC-O/L surface (concomitantly forced by non-solvent), and nitrogen bubbling promoted formation of uniform film at suspended material surface. Optimization of the water quantity was performed in a batch system in order to provide fulfillment of porous structure and formation water film at outer surface. Modification of NC surface with hydrophilic amino groups helps surface wetting and more importantly, the interaction of Lewis acid (iron ion) with Lewis base site (amino groups) provides initial nucleation center for magnetite precipitation. It was performed by continual bubbling of N2 at 90 °C, and by dropwise addition of an oxygen-free solution of 20 mL (0.86 g KNO3 and 5.95 g KOH) or 10 mL (0.38 g KNO3 and 2.65 g KOH) or for 20 min into dispersions of xylene/NC-MA/L and xylene/MC-O/L impregnated with appropriate quantity of FeSO4·7H2O, respectively. After precipitation, the reaction was heated for additional 60 min and cooled overnight. The adsorbent NC-MA/L-MG and MC-O/L-MG were isolated as black precipitate, washed with DW until getting free from sulfate ion in filtrate and dried. No sieving was necessary to apply for material obtained by this method. The multistep syntheses of NC-MA/L-MG and MC-O/L-MG adsorbents are illustrated in Fig. 1.

Modification paths applied for NC-MA/L-MG (a) and MC-O/L-MG (b) adsorbents synthesis.
Figure 1
Modification paths applied for NC-MA/L-MG (a) and MC-O/L-MG (b) adsorbents synthesis.

2.8

2.8 Characterization of adsorbents

Analysis of textural properties of obtained materials was performed by using Brunauer-Emmett-Teller (BET) method for the determination of the specific surface area, and the Barrett-Joyner-Halenda (BJH) method for the determination of the mesoporosity parameters. The X-ray diffraction (XRD) (BRUKER D8 ADVANCE) was used for phase and structural analysis of adsorbents. The surface morphology was examined by field emission gun scanning electron microscopy (FEG–SEM) with a field emission gun TESCAN MIRA3 XMU electron microscope and scanning electron microscope (SEM) by using JEOL JSM-6610LV. Distribution of nanoparticles on cellulose surface was further verified using transmission electron microscope (JEOL JEM-1400). Fourier-transform infrared spectra (FTIR) were collected on BOMEM (Hartmann & Braun) spectrometer and were employed to determine the surface groups and its interactions with arsenate. Raman spectra were collected with a XploRA Raman spectrometer from Horiba Jobin Yvon. The system employed laser at 532 nm (maximum output power 20–25 mW). The thermal stability of the samples was investigated by simultaneous non-isothermal thermo-gravimetric analysis (TGA) using a SETARAM SETSYS Evolution 1750 instrument. The measurements were conducted at a heating rate of 10 °C/min in argon or air atmosphere (flow rate was 20 mL/min) in the temperature range of 30–700 °C. Determination of amino group was performed according to the literature volumetric method and degree of amination obtained by Kaiser test (DAKaiser) (Vuković et al., 2011). The pH values at the point of zero charge (pHPZC) of the samples, i.e. the pH above which the total surface of the samples is negatively charged, were measured using the pH drift method (Erdem et al., 2004; Vuković et al., 2011). Quantification of the available amino group was determined by applying Kaiser test (DAKaiser – degree of amination obtained by Kaiser test (Tian et al., 2011), and by volumetric method (Vuković et al., 2011). The acid value (AV) was determined according to ASTM D3644 standard, and presented as mg KOH g−1 per gram of specimen.

The arsenic concentrations in the solutions after the adsorption and kinetic experiments were analyzed by inductively coupled plasma mass spectrometry (ICP-MS), using an Agilent 7500ce ICP-MS system (Waldbronn, Germany). ICP-MS detection limit was 0.030 μg L−1 and the relative standard deviation (RSD) of all arsenic species investigated was between 1.3 and 5.1%.

2.9

2.9 Batch adsorption experiments

Batch adsorption experiments were conducted in a 100 mg L−1 suspension with 10 mL of arsenate solution at pH 6.0 ± 0.1. Adsorption equilibrium and kinetic experiments were performed by varying initial arsenic concentrations: 0.10, 0.30, 0.50, 0.60, 1.0, 2.0, 3.0, 5.0, 7.0, 10.0, 15.0 and 20.0 mg L−1 at three temperatures of 298, 308 and 318 K. The initial pH (pHi) values were adjusted by addition of 0.5 mol L−1 HCl or 0.5 mol L−1 NaOH solutions. A laboratory pH meter, Mettler Toledo FE20/FG2 (Switzerland), with an accuracy of ±0.01 pH units, was used for the pH measurements. The systems were equilibrated at different rates of stirring, 100–150 rpm. The effect of contact time on adsorption kinetics was studied in the range 2–360 min at Ci = 100 μg L−1, and 90 min was selected as optimal time. Competitive kinetic study was conducted with 0.1 mg L−1 of silicate, chromate and phosphate solutions, contact time 90 min and m/V = 100 mg L−1. The sample solutions were filtered through a Millipore 0.22 μm membrane filter (Bedford, USA) and analyzed on Inductively Coupled Plasma Mass Spectrometry (ICP-MS). The arsenic uptake of adsorbents was evaluated by the use of mass balance Eq. (1):

(1)
q = C i - C f m V where q is adsorption capacity in mg g−1, Ci and Cf are initial and final arsenic concentrations in mg L−1, V is volume of the solution in dm3, and m is mass of adsorbent in g. All the experimental data presented were determined by using ICP-MS techniques, a given as mean values from three determinations. Results showed that the adsorption of arsenic on the container wall and filters was negligible.

The degree of conversion (degree of adsorption) could be expressed by a dimensionless number defined by Eq. (2):

(2)
α ( t ) = ( C i - C t ) / ( C i - C f ) where Ct is the concentration of arsenic at elapsed time of adsorption t, and this expression is used for calculation of the extent of adsorption.

To evaluate the regeneration capabilities, after adsorption experiments and washing with DW, wet adsorbents (m/V = 100 mg L−1) were redispersed in 20 mL of the solution used for desorption (NaOH/NaCl, 0.5/0.5 mol L−1). The amount of desorbed As(V) was measured after magnetic mixing for 3 h in a batch system. Five consecutive adsorption/desorption cycles were performed in triplicate.

2.10

2.10 Response surface methodology (RSM): experimental design of adsorbent preparation

Full data on applied RSM methodology are given in Supplementary data.

2.11

2.11 Statistical criteria used for the evaluation of the goodness of the modeling

Statistical methods applied for evaluation of the fitting are given in Supplementary data.

3

3 Results and discussion

3.1

3.1 Optimization of the adsorbents synthesis

In order to design effective, reusable and selective hybrid adsorbents for As removal an optimization study was conducted. Optimization goals were directed to syntheses of adsorbent with high capacity/efficiency, improved chemical and mechanical stability of MG deposit, and fast adsorption of arsenic at operational time scale. The optimization procedure, performed according to Sections 2.5 and 2.6, and according to experimental plan presented in Table S1, revealed that optimal performance was obtained by using the following:

  • 100 mL of FeSO4·7H2O (0.38 mol L−1) and 45 mL of pH adjustment solution (5.95 g KOH and 0.86 g KNO3) for NC-MA/L-MG adsorbent synthesis (Fig. 2(a)), and

  • 100 mL of FeSO4·7H2O (0.17 mol L−1) and 45 mL of pH adjustment solution (2.65 g KOH and 0.38 g KNO3) used for MC-O/L-MG adsorbent synthesis (Fig. 2(b)).

Contour diagram representing adsorbent capacity versus CFeSO4 and pH in the synthesis of NC-MA/L-MG (a), MC-O/L-MG (b), and capacity of both adsorbents vs amount of iron oxide (c).
Figure 2
Contour diagram representing adsorbent capacity versus CFeSO4 and pH in the synthesis of NC-MA/L-MG (a), MC-O/L-MG (b), and capacity of both adsorbents vs amount of iron oxide (c).

Results of the optimized procedure of adsorbent syntheses in a batch system (Sections 2.5 and 2.6) were applied, with an adjustment of the volume of used solution to provide reaction in a water thin/hydrophylic precursor material, for syntheses of NC-MA/L-MG and MC-O/L-MG adsorbent according to methodology described in Section 2.7. In that way, the highest adsorption capacities were obtained at ∼29% loaded iron oxide in NC-MA/L-MG and ∼13% in MC-O/L-MG (Fig. 2(c)).

The most probable structure of designed intermediates and adsorbents is presented in Fig. 1. Significance of the amino groups involved in the complexation was recently unequivocally presented (Cabaniss, 2011). Cations are strongly bonded to amino groups, by complexation/-chelation interactions, and higher nucleophilicity of amino groups, at pH > pHPZC, brings to stronger interaction with cation (Vuković et al., 2011). In that sense, it was of utmost interest to introduce number of amino groups through hierarchical branching of organic structure covalently bonded to NC and MC surface (Fig. 1). The total basic sites, 0.60 and 1.02 mmol g−1 found for MC-O/L and NC-MA/L, respectively, contributed to uniform/high MG loading. According to total basic sites determination, it was evaluated that ∼19 and ∼11 wt.% of introduced organic functionalities are present in NC-MA/L and MC-O/L, respectively. Modification of NC, with an acid value (AV) of 28, by MA to NC-MA, introduced ∼1.50 mmol g−1 of carboxylic group (AV = 150). Lower efficiency of NC-MA modification with EDA resulted in 0.92 mmol g−1 of amino group (∼61% conversion) in NC-MA/EDA. Despite that according to stoichiometric ratio considered for subsequent modifications of NC-MA/EDA to NC-MA/L (Fig. 1), 2.76 mmol g−1 of amino groups should be expected, and experimental value of 1.02 mmol g−1 was found. This indicates lower efficiency of modification introducing amino groups, present at NC-MA/EDA surface, due to steric hindrance of introduced voluminous tetraethyl ester of EDTA structure. Generally, optimal procedure for MG precipitation/complexation on NC-MA/L and MC-O/L was defined by using RSM methodology, was successful, according to Sections 2.5 and 2.6, but controlled precipitation and uniform distribution of MG on NC-MA/L and MC-O/L surface were achieved according to Section 2.7.

3.2

3.2 Characterization of the adsorbents

3.2.1

3.2.1 Textural properties and pHPZC

Multi-step synthesis of adsorbents was applied in order to design the material with an appropriate geometry, surface properties and pore structure. The textural properties and zero point of charge (pHPZC) before and after adsorption are summarized in Table 1.

Table 1 Textural properties, pHPZC of MG, NC-MA/L-MG and MC-O/L-MG adsorbent.
Adsorbent Specific surface area (m2 g−1) Pore volume (cm3 g−1) Pore diameter (nm) pHPZCa pHPZCb
MG 66.7 0.85 12.3 6.6 5.9
NC-MA/L-MG 85.3 0.71 15.2 6.5 5.7
MC-O/L-MG 35.5 0.45 18.4 6.1 5.8

pHPZC of NC was found to be <1, and pHPZC of NC-MA <3.

Before adsorption.
After adsorption.

Due to dimensional differences of parent materials, the higher value of specific surface area and pore volume was in favor of NC-MA/L-MG, while higher pore diameter was found for MC-O/L-MG compared to the raw MG. Literature data of unmodified nano- and micro fibrillated cellulose (Dufresne, 2012), showed that differences in surface area and pore diameter are generally influenced by the textural properties of substrate, applied modification method as well as post processing of obtained material. In contrast to mild thermal treatment applied for NC-MA/L-MG and MC-O/L-MG production, high temperature processing exerts significant change in textural properties of final material. The surface area and pore volume of thermally treated silica, after ligand immobilization, showed significant decrease in textural properties (Awual, 2015; Awual et al., 2016a, 2016b; Awual, 2016a, 2016b, 2016c). Regardless of obtained results both methodologies could find application in production of adsorbents of different properties. Precipitation of MG on NC-MA/L and MC-O/L provided such conditions applicable for preservation of organic branched structures with controllable deposition of nanoscale materials. The shift of pHPZC value (Table 1) indicates specific arsenate adsorption rather than electrostatic interaction, and formation of complexed/precipitated arsenic species at surface of magnetite based adsorbent is a main adsorption mechanism (Markovski et al., 2014a, 2014b).

3.2.2

3.2.2 XRD analysis

XRD diffractograms of NC and NC-MA, presented in Fig. 3(a), provide insight into structural changes in the course of NC modification. An increase in crystallinity of NC-MA was observed. Both diffractograms display two well-defined peak doublets around 2θ = 20° and 22.0° which correspond to (2 0 0) plane of cellulose crystal structure. Those doublets indicate the existence of cellulose I and cellulose II allomorphs (Peng et al., 2013). Modification with MA causes sharper signal, and increases contribution of cellulose II crystallites, and overall degree of crystallinity. In the sample of NC-MA signal at 12.24° corresponds to (−1 1 0) crystal plane of cellulose I (Klemm et al., 2011). The signal at 35.0° showed higher intensity for NC-MA, corresponds to (0 0 4) atomic plane of cellulose I (Peng et al., 2013). Also, it was shown that appearance of new crystalline peaks at 2θ ∼ 35° as well as at 41.0° indicates the increase in the crystallinity of NC-MA (Maiti et al., 2013).

XRD diffractograms of (a) NC and NC-MA, and (b) MC-O/L-MG and NC-MA/L-MG.
Figure 3
XRD diffractograms of (a) NC and NC-MA, and (b) MC-O/L-MG and NC-MA/L-MG.

Phase and structural analysis of NC-MA/L-MG and MC-O/L-MG confirmed the presence of MG (Fe3O4), as it was indicated by peak at 30.16, 35.52, 37.16, 43.17, 54.22, 57.10, 62.70° (JCPDS 19-629) (Fig. 3(b)). Differences in peak broadness and intensity indicate lower degree of crystalline phase of embedded MG in MC-O/L-MG. The mean crystallite size, calculated by using Scherrer equation based on the half width of the most intense peaks in XRD patterns, indicates 22 and 19 nm of grain sizes of MG present on NC-MA/L-MG and MC-O/L-MG surface, respectively. These values are somewhat lower than ones observed by transmission electron microscopy (TEM) (Fig. 7). A somewhat lower crystallite size, found for MC-O/L-MG, is due to lower number of amino groups (crystallization centers) accessible for iron coordination. Obtained results indicated the high coverage of NC-MA/L and somewhat lower for MC-O/L by MG. Due to the known fact that MG and maghemite (γ-Fe2O3) have similar XRD pattern it was necessary to apply Raman spectroscopy in order to clearly distinguish nanocrystalline form of deposited iron oxide.

3.2.3

3.2.3 Raman spectroscopy

Raman spectra of the NC, NC-MA/L and NC-MA/L-MG, and MC and MC-O are shown in Fig. 4. The region below 1750 cm−1, observed at Raman spectra of NC and NC-MA/L, originates from the conformation of cellulose skeleton, while the region above 2888 cm−1 and higher than 3000 cm−1 originates from hydrogen bonding of OH group (Szymańska-Chargot et al., 2011). Comparing the Raman spectra of the NC and NC-MA/L, the overlapped bands, observed in the region 3215–3405 cm−1, are due to OH and NH2 stretching vibration (Gabelica, 1976). Also, the signals at 1456 cm−1 and 1473 cm−1, a characteristic for methylene bending modes of two stereo-chemically nonequivalent CH2OH groups of the crystalline NC skeletons, merge into one single signal at 1456 cm−1 in NC-MA/L. This change explained the differences between the vibrational behavior of crystalline and amorphous domain in NC (Schenzel et al., 2005), and indicated transition from crystalline to amorphous form of these two CH2 bending modes during surface modification (Schenzel and Fischer, 2001; Wiley and Atalla, 1987). Also, the intensity of the characteristic peak of highly crystalline NC (1473 cm−1), and other signals decrease in comparison with amorphous NC-MA/L domain.

Raman spectra of (a) NC, NC-MA/L and NC-MA/L-MG, and (b) MC and MC-O.
Figure 4
Raman spectra of (a) NC, NC-MA/L and NC-MA/L-MG, and (b) MC and MC-O.

Weak bands, observed in region from 964 cm−1 to 1113 cm−1, originate from C6 atom bending vibrations (H—C—C and H—C—O), while the strong stretching vibration of unhydroglucose skeleton is observed at 1145 cm−1 (C—C, C—O and C—O—C) (Schenzel et al., 2005; Wiley and Atalla, 1987). The NC backbone deformation vibration bands, δ(CH2), δ(HCC), δ(HCO) and δ(COH), are observed at 1374 cm−1 (Zhang et al., 2011). Wide bands observed in the region between 1565 cm−1–1992 cm−1 and 900 cm−1–1500 cm−1 originate from the MA and EDTA moieties in NC-MA/L, respectively.

The non-polarized spectrum of NC-MA/L-MG shows magnetite predicted phonon bands as characteristic peak at 670 cm−1 (Shebanova and Lazor, 2003). Due to susceptibility of MG to transform into maghemite, under exposure to laser, two weak peaks appeared at 350 and 500 cm−1. The absence of NC characteristic bands in NC-MA/L-MG spectrum indicates homogenous distribution/entire surface coverage by deposited MG on amino functionalized NC.

The Raman spectra of the MC and MC-O showed similar peak structure and position, except that cellulosic structure of MC basis participates at higher extent in response to laser impulse. Regardless of adjustment of laser power the Raman spectrum of MC-O/L-MG sample was not obtained.

3.2.4

3.2.4 FTIR spectroscopy

The FTIR spectra of intermediates and adsorbents, before and after arsenate adsorption, are presented in Fig. 5. From the FTIR spectra of all intermediates, a broad band about 3423 cm−1 ascribed to OH and NH stretching vibrations can be observed. Adsorption of arsenate oxyanions causes shifting of the stretching vibration of OH band to 3451 cm−1. Differences in the peak intensity, peak shifting and appearance of peak after adsorption indicate arsenate bonding by appropriate functional groups. Carbonyl group absorption at 1714 cm−1, found in NC-MA, is missing in the spectrum after MG modification. Absorption peak, at 1635 cm−1, originates from deformation vibrations of OH bond, and is negligibly shifted to 1637 cm−1 after MG modification. Broad bands at ≈1714 and ≈1635 cm−1, are shifted to ≈1637 cm−1, in NC-MA/EDA and NC-MA/L, a band corresponding to amide stretching vibration (amide I) coupled with in-plane N—H deformation vibration. The bands in the region 750–1000 cm−1, and one at 1373 cm−1, show the presence of sulfonates and amino groups (Socrates, 2004). The strong bands at 582 cm−1, assigned to Fe—OH vibrations present at MG surface, after arsenate adsorption appeared at 617 cm−1. This region corresponds to vibration of As—O—Fe complexed/precipitated arsenate species, and at different surface coverages, bidentate binuclear and mononuclear monodentate complex could be formed, where As—OH free bonds are present as unprotonated and/or protonated depending on pH (Markovski et al., 2014a, 2014b; Grossl et al., 1997).

FTIR spectra of (a) NC, NC-MA, NC-MA/EDA and NC-MA/L and (b) NC-MA/L-MG, MC-O/L-MG, NC-MA/L-MG/As and MC-O/L-MG/As.
Figure 5
FTIR spectra of (a) NC, NC-MA, NC-MA/EDA and NC-MA/L and (b) NC-MA/L-MG, MC-O/L-MG, NC-MA/L-MG/As and MC-O/L-MG/As.

3.2.5

3.2.5 Thermal analysis (TGA)

The success of nano/microcellulose modification and thermal stability of composites constituent was evaluated by the use of TGA. TGA and differential thermal analysis (DTG) curves, recorded for NC, MC, NC-MA/L, NC-MA/L-MG and MC-O/L-MG, are shown in Fig. 6. The TGA curve and DTG peaks, observed at 258 °C for NC and 260.5 °C for MC, indicate similar thermal behavior of both materials in whole temperature range. As it can be observed, the thermal decomposition of NC and MC occurred in three steps: (i) 30–225 °C, (ii) 225–325 °C, and (iii) 325–700 °C range. In the first stage, ∼5% and ∼11% weight loss were observed due to dehydration of NC and MC, respectively. In the second stage, almost 61% mass loss occurred due to the splitting/thermal decomposition of cellulose structure, i.e. chain scission which evolves gaseous products leaving condensation/carbonaceous residues. After 325 °C, 20% and 18% weight loss were observed due to the degradation of the NC and MC condensed carbonaceous material, respectively. The end of first region is shifted to the lower temperature (145 °C) for the NC-MA/L sample. Also, the weight loss difference, ∼12%, between NC and NC-MA/L is similar to evaluated organic fraction, obtained according to amino groups calculation ∼18% (Section 3.1). DTG peak for NC-MA/L occured at 158 °C refers to dehydration and evaporation of residual solvents, while DTG peak at 273.4 °C refers to cellulose backbone decomposition.

TGA (a) and DTG (b) curves of NC, MC, NC-MA/L, NC-MA/L-MG and MC-O/L-MG.
Figure 6
TGA (a) and DTG (b) curves of NC, MC, NC-MA/L, NC-MA/L-MG and MC-O/L-MG.
FEG-SEM images of NC-MA/L (a), NC-MA/L-MG (b), MC-O/L-MG (d) and (e); and TEM images of NC-MA/L-MG (c) and MC-O/L-MG (f).
Figure 7
FEG-SEM images of NC-MA/L (a), NC-MA/L-MG (b), MC-O/L-MG (d) and (e); and TEM images of NC-MA/L-MG (c) and MC-O/L-MG (f).

The higher stability, comparing to NC and MC, and similar thermal behavior of NC-MA/L-MG and MC-O/L-MG were noticed (Fig. 6(a)). Higher thermal stability is due to contribution of MG protective layer which forms, by sintering of MG nanocrystallite, a stable core shell-like structure. The first stage of mass loss (<300 °C) is due to removal of physically adsorbed water, residual reactants and amorphous cellulose domain. The weight loss in the range from 300 to 450 °C is ascribed to the rapture of weak head to head acetal linkage between cellobiose units in an amorphous/lower crystallinity domain of cellulose. Corresponding DTG peaks could be observed at 306.8 °C, 324.6 °C and 379 °C for MC-O/L-MG and 394.2 °C for NC-L-MG, respectively. The third stage, at >400 °C, takes place by random scission/gradual decomposition of main cellulose polymeric chain. A residual inorganic material of 42% and 28% in an inert atmosphere, and comparative experiments in air showed that 33% and 17% residuals in NC-L-MG and MC-O/L-MG were obtained, respectively. On the DTG curves of NC-L-MG and MC-O/L-MG peaks observed at temperatures >470 °C refer to thermal oxidation and dehydroxylation of precipitated magnetite (Przepiera and Przepiera, 2001).

3.2.6

3.2.6 Morphological characterization

The morphology of the synthesized adsorbents was analyzed by using SEM/FEG–SEM techniques, results given in Fig. 7(a), (b), (d) and (e), while TEM results are given in Fig. 7(c) and (f).

Morphology of NC-MA/L (Fig. 7a) indicated surface coverage by organic material due to chemical modification, and subsequent MG deposition affects the morphological structure of the surface of NC-MA/L-MG (Fig. 7b). SEM micrograph of NC-MA/L-MG adsorbent obtained at lower magnification is shown in Fig. S1 (Supplementary material).

SEM micrograph of MC-O/L-MG adsorbent is given in Fig. 7(d) and (e). The evaluated average diameter of MC, MC-O and MC-O/L-MG was in the range 30 ± 1.2 μm. The modified MC fibers showed small diameter changes of parent cylindrical strands due to organic modification and MG precipitation at MC-O/L surface. Evolution of the MC-O/L-MG morphology can be explained by heterogeneous reaction which takes place by MG precipitation on terminal amino branched structure with incorporated NC grains on the fiber surface.

TEM analysis (Fig. 7(c) and (f)) confirmed nanoscale dimension and irregular distribution of cubic MG nanoparticles onto irregular/porous morphology of NC-MA/L and MC-O/L. It can be concluded that morphology of NC-MA/L-MG and MC-O/L-MG adsorbents is the result of both physical structure of nano/microscale material, and chemical structure/properties of functional groups present at NC-MA/L and MC-O/L surface.

3.3

3.3 Adsorbents performance in non-competitive conditions

3.3.1

3.3.1 pH-dependent, equilibrium, thermodynamic and competitive adsorption study

Considering significance of pH influences on arsenic speciation and ionization state of adsorbent surface, it was important to study influence of pH (in the range 3–10) on effectiveness of As removal. The results of the pH-dependent adsorption study, i.e. arsenate removal versus initial pH (pHi) for both NC-MA/L-MG and MC-O/L-MG adsorbents, and change of final pH (pHf) in relation to initial pH (pHi) are shown in Fig. 8(a). The extent of adsorption pHf/pHi relation is affected by both release/consumption of hydrogen ions, i.e. protonation/deprotonation of MG surface, and pH-dependent arsenic speciation. These factors determine that effective As(V) adsorption was obtained in pH range 3.0–7.0 (Fig. 8(a)), while pHf/pHi relationship indicates similar arsenic adsorption mechanism with respect to both adsorbents. Similar behavior was found for As(III) pH-dependent adsorption (Fig. S2).

Effect of pH on adsorption of As(V) on NC-MA/L-MG and MC-O/L-MG, and pHf versus pHi dependence (a); Freundlich fitting of adsorption data for NC-MA/L-MG (b) and MC-O/L-MG adsorbents (c).
Figure 8
Effect of pH on adsorption of As(V) on NC-MA/L-MG and MC-O/L-MG, and pHf versus pHi dependence (a); Freundlich fitting of adsorption data for NC-MA/L-MG (b) and MC-O/L-MG adsorbents (c).

The pKa values of arsenic acid: pKa1 ∼ 2.3, pKa2 ∼ 7.0 and pKa3 ∼ 11.5, indicate that molecular form mainly exists in a solution at pH < 2.0, and pH-dependent equilibrium of anionic species, H2AsO4 or HAsO42−, exists at pH in the range 2–10 (Fig. S3) (Issa et al., 2011). At pHi < pHPZC (Table 1), the protonation of surface functional generates positive charge which contributes to favorable bonding of negatively charged arsenate ions (Fig. S3). Both surface state and As speciation play significant contribution to electrostatic interaction (attraction/repulsion) between surface/ions causing intensity of As flux toward the specific adsorption sites. Enhancement of electrostatic attraction is thus highly feasible for As(V) species, while it is of minor importance for the neutral form of arsenous acid (Fig. S3). At pHi > pHPZC, deprotonation of surface functional groups, increased concentration of negatively charged arsenic species and competition of hydroxyl ion at higher pH contributed to enhanced repulsion at the boundary layer of the adsorbent interface, i.e. lower adsorption efficiency is a consequence.

Simultaneous determination of arsenic and iron by ICP-MS showed that at pH < 5 significant dissolution of MG precipitate was noticed. Good adsorption capacity and noticeable iron dissolution at pH < 5 dictated that selection of pH 6.0 was the best choice, and this pH was used in subsequent adsorption experiments.

The state of adsorption equilibrium was analyzed by fitting experimental data with various isotherm models (Foo and Hameed, 2010) by applying nonlinear least squares fitting to adsorption data. Statistical criteria (Section 2.11) were used to evaluate the model fitting of adsorption data. Results indicated that the good fitting was obtained by using Langmuir, Eq. (S2), and Freundlich, Eq. (S3), isotherm models.

Adsorption results presented in Fig. 8(b) and (c) and Table 2 showed that good fitting of the experimental data with Freundlich isotherm indicates multilayer adsorption on heterogeneous surfaces. Significantly higher adsorption capacity for NC-MA/L-MG (85.3 mg g−1), in comparison with MC-O/L-MG (18.5 mg g−1), was obtained by using Langmuir model (Table S2). The favorable textural parameters of NC-MA/L-MG contribute to higher arsenate uptake, i.e. larger number of available adsorptions sites at nanocomposite surface (Table 1). Somewhat lower capacity was obtained with NC-MA/L-MG and MC-O/L-MG in a processes of As(III) removal (Table S3), 68.2 and 17.8 mg g−1, respectively. This result is indicated on a wider applicability of synthesized adsorbents. Also, comparative adsorption study of arsenic removal by using NC-MA/L showed that 15.4 mg g−1 and 4.8 mg g−1 were found for As(V) and As(III) at pH 6.0 ± 0.1, respectively. Even lower capacities were obtained for MC-O/L (<6 mg g−1 for As(V) removal). These results clearly indicate that validity of the modification of NC-MA/L and MC-O/L with MG was justified by production of high performance NC-MA/L-MG and MC-O/L-MG adsorbents. Thermodynamic parameters are calculated, according to Eqs. (S4) and (S5) (Vuković et al., 2011), based on adsorption studies performed at different temperatures and calculated KL Langmuir constant (Table S2), and results are given in Table 2.

Table 2 Adsorption isotherm (Freundlich) and thermodynamic parameters for arsenate removal.
T (K) Kf ((mol/g)(L/mol)1/n) 1/n R2
Freundlich parameters
MG 298 0.0539 ± 0.0016 0.396 ± 0.012 0.997
308 0.0473 ± 0.0014 0.480 ± 0.014 0.995
318 0.0357 ± 0.0012 0.349 ± 0.011 0.993
NC-MA/L-MG 298 0.2412 ± 0.0961 0.562 ± 0.024 0.995
308 0.2005 ± 0.0802 0.536 ± 0.022 0.994
318 0.1474 ± 0.0598 0.503 ± 0.020 0.993
MC-O/L-MG 298 0.0032 ± 0.00008 0.188 ± 0.0047 0.995
308 0.0025 ± 0.00006 0.258 ± 0.0065 0.997
318 0.0020 ± 0.00005 0.229 ± 0.0057 0.994
T (K) ΔG° (kJ mol−1) ΔH° (kJ mol−1) ΔS° (J mol−1 K−1) R2
Thermodynamic parameters
MG 298 −40.56 ± 0.81 4.31 ± 0.08 150.48 ± 3.01 0.992
308 −42.05 ± 0.82
318 −43.57 ± 0.87
NC-MA/L-MG 298 −38.66 ± 1.16 10.77 ± 0.32 165.69 ± 4.97 0.978
308 −40.23 ± 1.21
318 −41.98 ± 1.26
MC-O/L-MG 298 −39.40 ± 0.985 10.99 ± 0.27 168.80 ± 4.22 0.954
308 −40.90 ± 1.022
318 −42.78 ± 1.069

The negative value of Gibbs free energy together with positive enthalpy and entropy changes indicates spontaneity and endothermicity of adsorption processes (Table 2). The positive ΔH° confirms favorable adsorption at higher temperatures. It means that disruption of water hydration shells and transport of arsenic anionic species through the bulk solution, within the pores and through the surface boundary layer are more intensive process at higher temperatures. The transport of exchangeable arsenic oxyanions to adsorption site, and number of (exchangeable) water molecules were released into bulk solution contributing to the increase in entropy change (Markovski et al., 2014a, 2014b). At steady-state condition randomness at adsorbent/solution interface is increased due to different intermolecular interactions which contribute to positive entropy change, i.e. adsorption is an entropy-driven process. All of these elementary processes during equilibration contribute to positive enthalpy change. Thermodynamic parameters of both As(V) (Table 2) and As(III) (Table S3) were similar to ones given by Liu et al. (2015).

Preliminary evaluation of the potential of use of both adsorbents NC-MA/L-MG and MC-O/L-MG for arsenic removal from real water samples, contaminated natural water from the area of the city of Zrenjanin was used (located in Vojvodina, Serbia). Sample used in this work was collected at the point before any purification treatment. Total arsenic concentration, determined by ICP-MS, was found to be 106 μg L−1. Other ionic species of interest present in real water sample are shown in Table 3.

Table 3 Content of ionsa in real water samples (mg L−1).
Ions Concentration in a real water
PO43− 5.6
SO42− 42.5
Ca2+ 15.3
Mg2+ 9.1
SiO44− 5
Cl 1.2
HCO3 856
pH of the sample was 7.4 and it was adjusted to pH 6 ± 0.1.

The total arsenic concentration includes following arsenic species: inorganic, As(III) and As(V), and methylated organic arsenic species (monomethylarsenic acid, MMAs(V), and dimethylarsenic acid, DMAs(V)). Due to this it was necessary to perform preliminary determination of present arsenic species. In a previous report we developed simple and efficient method for separation and determination of inorganic arsenic (iAs) and organic arsenic (oAs) in natural water (Issa et al., 2011). Two types of resins were used for determination of arsenic species: a strong base anion exchange (SBAE) resin and HY–AgCl based on the activity of silver-chloride loaded on SBAE resin. The SBAE resin has been chosen as a selective resin for the separation of As(V) from As(III), and HY–AgCl for quantitative separation of molecular and ionic forms of iAs and oAs. Molecular form of As(III) that exists in the water at pH < 8.0 was not bonded with SBAE, which was convenient for direct determination of As(III) concentration in the effluent. Experiments performed in a column system filled with HY–AgCl showed that oAs species are not presented in a sample water used. Additional experiment performed in a column system filled with SBAE resin showed that As(III) was present at the level less than 5 μg L−1. In order to perform As(III) oxidation air bubbling for two hours was performed, and determined As(V) in oxidized solution was found to be 104 μg L−1.

Preliminary adsorption study performed with real water sample (Table 3), spiked with 10 mg L−1 of As(V), showed 19% decrease in adsorption capacity for NC-MA/L-MG (69.1 mg g−1), and 8% for MC-O/L-MG (17.0 mg g−1) with respect to experiments performed under non-competitive condition. These results showed possible application of synthesized adsorbents, even in the presence of significant concentration of interfering ions in natural water, in a processes of As(V) removal from groundwater.

3.3.2

3.3.2 Adsorption kinetics in a non-competitive condition

Due to complexity of adsorption processes, i.e. contribution of mass transfer and chemical reaction, different kinetic models were used to fit experimental data: pseudo-first order, pseudo-second order (PSO), Roginsky-Zeldovich-Elovich, and second-order rate equation. Also diffusion models: liquid film linear driving force rate equation, liquid film diffusion mass transfer rate equation, homogeneous solid diffusion model, Weber-Morris model (WM), Dunwald-Wagner model, and double exponential model (Markovski et al., 2014a, 2014b) were used. PSO and W–M models showed the best fitting (Fig. 9 and Table 4). Additionally, result of comparative kinetic experiments with As(III), and PSO fitting showed higher rate constant, 0.762 ± 0.022 g mg−1 min−1 for MC-O/L-MG at 25 °C, which is an additional evidence of high performances of obtained material.

Plot of arsenate adsorption on MG by using PSO (a) and WM (b) model at 25 °C (m/V = 100 mg L−1, Ci = 0.1 mg L−1, pH 6.0 ± 0.1), and diagram of activation energies Ea vs α c).
Figure 9
Plot of arsenate adsorption on MG by using PSO (a) and WM (b) model at 25 °C (m/V= 100 mg L−1, Ci = 0.1 mg L−1, pH 6.0 ± 0.1), and diagram of activation energies Ea vs α c).
Table 4 Kinetic and activation parameters of As(V) removal by using MG, NC-MA/L-MG and MC-O/L-MG adsorbent.
Adsorbent Pseudo-second-order model Weber–Morris model
qe (mg g−1) K2a R2 kp1b C (mg g−1) R2 kp2b R2 kp3b R2
MG 1.047 ± 0.024 0.156 ± 0.017 0.985 0.151 ± 0.014 0.110 ± 0.047 0.957 0.0096 ± 0.0035 0.807
NC-MA/L-MG 1.0463 ± 0.0084 0.189 ± 0.008 0.997 0.204 ± 0.011 0.031 ± 0.030 0.987 0.0181 ± 0.0047 0.829
MC-O/L-MG 1.0155 ± 0.0095 0.541 ± 0.033 0.989 0.320 ± 0.063 0.065 ± 0.115 0.926 0.0827 ± 0.0025 0.998 0.016 ± 0.004 0.813
Ea (kJ mol−1) ΔH25# (kJ mol−1) ΔG25# (kJ mol−1) ΔS25# (J mol−1 K−1)
Activation parameters
MG 20.10 ± 0.96 17.63 ± 0.86 77.23 ± 3.8 −200.8 ± 9.36
NC-MA/L-MG 13.28 ± 0.65 10.8 ± 0.54 76.96 ± 3.7 −222.5 ± 10.6
MC-O/L-MG 10.87 ± 0.53 8.39 ± 0.42 74.55 ± 3.5 −221.9 ± 10.3
g mg−1 min−1.
mg g−1 min−0.5.

The results obtained following the W–M model showed three successive linear steps for MC-O/L-MG, while two-line correlation lines were obtained for NC-MA/L-MG (Fig. 9). Fast kinetics in the first step was followed by medium adsorption rate in the second, and very slow attainment of equilibrium in the third step was found for MC-O/L-MG. Higher values of the rate constant, found for MC-O/L-MG (Table 4), reflect better diffusibility/easier accessibility of surface active sites by sorbate due to shortening the path of ions inside the adsorbent body.

It is well known that multi-step nature of adsorption processes could be affected by bulk diffusion, boundary layer diffusion, intra-particle diffusion and surface adsorption. The first linear steep part demonstrates external mass transfer from bulk solution to the most available adsorptive sites at outer adsorbent surface. The second and the third parts of the adsorption process significantly depend on adsorbent porosity, i.e. pore structure, geometry and network density. Due to the concentration gradient, the ions diffuse through bulk solution and treelike pore system extending into adsorbent interior to reach all available surface active sites (Sun et al., 2014; Yousif et al., 2016). The intra-particle resistance slows down adsorbate transport, i.e. net transport in a direction of variable time-dependent concentration gradient. At the final stage, the adsorption takes place at low rate until the saturation of all available surface sites is achieved.

The activation parameters: entropy of activation ΔS#, Gibbs free energy of activation ΔG# and enthalpy of activation ΔH# were calculated from kinetic data, obtained at 25, 35, and 45 °C, by using Eqs. (S6)–(S9). The results are presented in Table 4.

The activation energies of 13.28 and 10.87 kJ mol−1 for As(V) removal with NC-MA/L-MG and MC-O/L-MG, respectively, were obtained. These results are in agreement with similar result obtained for synthetic goethite (Lakshmipathiraj et al., 2006). Adsorbate transport is generally controlled by diffusive mass transfer, and thus rates of approaching to equilibrium usually increase with increase in temperature. The results show that the intra-particle diffusion is a rate-controlling step since the activation energy is low and within the range of 8–22 kJ mol−1, characteristic value for diffusion-controlled processes such as ion-exchange/complexation (Haring, 1942).

The improvement of the analysis of kinetic data for whole period of adsorption was performed by applying model free kinetics of adsorption, i.e. determination of activation parameters by iso-conversional method. Model free kinetics was selected in order to get the most reliable system responses at appropriate period of time, i.e. at the beginning, middle and plateau (approaching to steady state condition) parts of the process. The adsorption rate can be described as the amount of adsorbed pollutant per unit of time by using dimensionless number α. It denotes time-dependent fraction of bounded arsenic, and could be described by Eqs. (S10)–(S13). A plot of −ln t versus 1/T at appropriates α yield Ea from the slope regardless of the kinetic/adsorption model was used. The reason for choosing the ‘model free’ method of kinetic analysis is based on the fact that any misidentification of the kinetic model has got a profound effect on the values obtained for the Arrhenius parameters. Sewry and Brown (2002) and Galwey (2003) have critically reviewed the use of various kinetic models and asserted that caution has to be exercised in choosing the right kinetic model. By applying the fitting of conversion-time data, using predetermined point for calculation of α value, obtained results are presented in Fig. 9(c).

From the slope of correlations Eq. (S12), at appropriate T and α, a series of activation energies data (Ea) were calculated, and results are presented in Fig. 9(c). The plots show that overall adsorption mechanism progressively changed from the initial point of the process to attainment of equilibrium. In general, Ea varies from 15 to 18 kJ mol−1 for NC-MA/L-MG, and from 12 to 14 kJ mol−1 for MC-O/L-MG. In the first part of adsorption (0 < α < 0.3), the favorable process was noticed for MG due to synergetic effect of geometry/surface group density. Afterward, a steady region (0.3 < α < 0.7) was attained, and followed by steep increases in the course of equilibration. Similar trend of low change of Ea, for both NC-MA/L-MG and MC-O/L-MG, was found for α < 0.7, and α > 0.7, and somewhat higher slope was found for NC-MA/L-MG. Generally, the change of Ea values (Fig. 9) indicates that adsorption is diffusion controlled process, and it is in accordance with the concept of variation of Ea in a real system due to multi-step nature of most processes encountered in practice. The time-dependent change in kinetics is attributed to textural properties, material structure/defects at exposed plane, steric hindrances and electrostatic attraction/repulsion between adsorbed/released molecules/ionic species. The significance of external/internal mass transport and diffusion inside the pores was indicated (Markovski et al., 2014a, 2014b). The obtained activation energy for the As(V) removal with vacuum and freeze dried goethite modified macroporous resin was 18.63 and 20.60 kJ mol−1 (Taleb et al., 2015), which indicates that similar phenomena dictate diffusional transport inside morphologically/chemically similar porous system.

3.3.3

3.3.3 Comparison of kinetic/adsorption data of magnetite based adsorbent

Additionally, performance of synthesized adsorbents was evaluated from comparison with other magnetite based adsorbents. Among the large numbers of cellulose based adsorbents synthesized up to date, most of them are claimed and considered to be the best candidates to feasible and applicable for effective arsenic removal. The adequate properties of cellulose based adsorbents, with respect to established criteria in the areas of environmental protection, mainly arise from their low cost and adsorption performances. The literature survey, shown in Table 5, of the data on adsorption kinetics and maximum adsorption capacity of the materials of interest is presented.

Table 5 Adsorption capacity and kinetic data of As(V) removal by using magnetite based/other adsorbents.
Adsorbent K2 (g mg−1 min−1) Adsorption capacity (mg g−1) References
Magnetite based adsorbent (batch study)
Magnetite 0.156 91.2 This paper
Magnetite 11.33 mg/min 88.4 Lin et al. (2012)
Magnetite/Activated carbon fiber 0.0081 4.16 Zhang et al. (2010)
Magnetite/Starch 248 An et al. (2011)
NC-MA/L-MG 0.189 85.3 This paper
Cellulose Cu 0.0187 98.9 Yousif et al. (2016)
Amino/G3MA/TiO2/Cellulose 0.010 108.7 Anirudhan et al. (2012)
Amino/Dianhydride/Cellulose 0.64 12.06 Singh et al. (2015)
Amino/Cellulose 0.029 187 Anirudhan and Jalajamony (2010)
MC-O/NC-L-MG 0.540 18.5 This paper
Magnetite/Cellulose fiber 0.00005 2.46 Hokkanen et al. (2015)
PDMAEMA/Cellulose fiber 27.9 Tian et al. (2011)
Amino/Cellulose fiber 0.0011 75.1 Yu et al. (2013)
Other adsorbents (column study)
Zirconium(IV)-loaded monophosphonic acid resin 0.0029 mmol/ml Awual et al. (2011)
Zirconium(IV) loaded fibrous adsorbent 0.127 mmol/g Awual et al. (2012)
Crosslinked poly(allylamine) resin 3.5 mmol/g Awual and Jyo (2009)
Commercial weak-base resins (Diaion WA20) 0.239 mmol/g Awual et al. (2013)
Commercial weak-base resins (Diaion WA30) 0.217 mmol/g Awual et al. (2013)
Weak-base anion exchange fiber 0.941 mmol/g Awual et al. (2008)

Bold values represent results from this work.

The diversity of experimental condition makes impossible convenient overview/comparison of the literature and presented adsorption data. General consideration of the relation of adsorbent performances/textural properties indicates positive dependence of increased surface area with respect to obtained adsorption capacity. According to the results presented (Table 5), the qmax was found in a wide range between 2.46 mg g−1 for Magnetite/Cellulose fiber and 108.7 mg g−1 for Amino/G3MA/TiO2/Cellulose. The qmax obtained from the Langmuir model for the NC-MA/L-MG was found to be 85.3 mg g−1 at 25 °C, and lower for the MC-O/L-MG, 18.5 mg g−1 was found, which is comparable to the literature data. It means that synthesized high performance adsorbents offer an acceptable alternative to those published in the literature or eventually used in a real processes, and production of adsorbents based on presented material offers many benefits to existing ones. Fast attainment of adsorption equilibrium, i.e., kinetic rate of arsenic adsorption in the presence of the MC-O/L-MG mainly exceeds the values found in the literature (Table 5), and capacity of NC-MA/L-MG is similar to those found in the literature. Obtained results are in accordance with general experienced rules on relationship between textural data and adsorbent performance, i.e. kinetic and adsorption data. The developed fibers integrated mechanical integrity (stability), hydraulic properties and arsenic selectivity. Despite lower adsorption capacity, another factor contributes to substantial improvement of its performance, and namely the results of adsorption kinetics open an alternative for the use of MC-O/L-MG exchange fibers relative to NC-MA/L-MG counterparts. The results presented point out to the advantage of application of micro-sized adsorbent which contributes to adsorption performances of chemically identical but physically different materials NC-MA/L-MG and MC-O/L-MG.

3.3.4

3.3.4 Competitive kinetic study

Except high capacity and fast adsorption rate of arsenic removal, the synthesized adsorbents should possess high affinity/selectivity with respect to selected pollutant. Similar to non-competitive kinetic study, the data obtained in competitive study of arsenate removal in the presence of single interfering ion: SiO44−, CrO42− and PO43− ion were processed by using PSO equation. The other kinetic models were of limited validity to be applied for description of the competitive kinetic study. Therefore, PSO model was used and results of fitting, obtained from plots qt − t and t/qt − t, are given in Table 6.

Table 6 Results of the PSO fitting of As(V) removal in the presence of single interfering ions.
Non-competitive Competitive condition
SiO44−a CrO42−a PO43−a
MG qeb 1.047(±0.024) 0.998(±0.014) 0.938(±0.014) 0.926(±0.016)
K2c 0.156(±0.017) 0.144(±0.045) 0.141(±0.040) 0.138(±0.037)
R2 0.985 0.989 0.990 0.992
NC-MA/L-MG qe 1.046(±0.084) 0.999(±0.013) 0.936(±0.019) 0.918(±0.028)
K2 0.189(±0.008) 0.181(± 0.022) 0.179(±0.020) 0.140(±0.027)
R2 0.997 0.987 0.989 0.991
MC-O/L-MG qe 1.016(±0.010) 0.968(±0.015) 0.932(±0.032) 0.886(±0.029)
K2 0.541(±0.033) 0.536(±0.029) 0.519(±0.022) 0.489(±0.0)
R2 0.989 0.990 0.982 0.992
Ci = 0.1 mg L−1, m/V = 100 mg L−1, t = 90 min.
mg g−1.
g mg−1 min−1.

On the other hand, calculated values show good agreement with the experimental data for PSO kinetics which means that sorption depends on the both concentration of arsenic and adsorbent surface functionalities. Single competitive As(V) adsorption showed the highest influence of phosphate (11.5% decrease of K2) for MG, 25.9% for NC-MA/L-MG and 9.6% for MC-O/L-MG. Similar trend but of lower extent was found for chromate and the lowest influence of silicate was noticed. Largest adverse effect showed phosphate due to similar chemistry with arsenate. The acid dissociation constants of phosphoric acid are pKa1 = 2.16, pKa2 = 7.21 and pKa3 = 12.32, which indicates that at pH ∼ 6 monovalent phosphate ion, H2PO4, is a dominant competitive species (Awual et al., 2011). Selection of pH 6 in adsorption study was optimal due to an increase in pH, hydroxide concentration increases so that interference of hydroxide with the hardest Lewis acids anion in water increases resulting in decrease in arsenate and phosphate adsorption and phosphate showed lower adsorption capacity at pH 6 than at higher pH on magnetite nanoparticles (Abo Markeb et al., 2016). Among the other interfering ions commonly present in natural water, sulfate, as divalent ions, shows greater competition through enhanced electrostatic interaction. Selection of pH 6, i.e. pH ≈ pHPZC, contributes to lower extent of electrostatic interaction of all ions with MG surface. Therefore, the main benefit of the designed system adsorbent/adsorption condition arises from achievement of desired levels of arsenate removal in the presence of interfering ions. Silicate showed low effect which was less than 8%.

Except this, experiments performed with NC-MA/L-MG in the presence of different concentrations of KCl, KNO3 (1 and 5 mg L−1), arsenic removal were slightly enhanced. Due to bonding of negatively charged arsenic species, concentration of negative charge is increased in a electrical double layer (Grossl et al., 1997), and at higher ionic strength of solution, increased concentration of counter cations could compensate the surface negative charges. Observed fact that increased ionic strength causes increased arsenate uptake indicating that formation of inner-sphere complexes (Markovski et al., 2014a, 2014b) is the main adsorption mechanism. Additionally, the low influence of interfering ions (Table 6) and pHPZC decreases after adsorption (Table 1) are the properties indicative for the formation of inner-sphere surface complexes. Ligand exchange is a final/crucial step at surface which occurs through formation of monodentate complex, while the next slow step represents a second ligand exchange resulting in the formation of an inner-sphere bidentate binuclear complex (Grossl et al., 1997). Formation of inner-sphere complexes is a process of highest probability, and other types of interaction between As species and protonated hydroxyl groups by forming outer-sphere surface complexation could be operative at lower pH (Liu et al., 2015).

Additionally, competitive kinetic experiments, performed with NC-MA/L-MG and MC-O/L-MG and natural water sample (Table 3), were carried out to find out the level of the efficiency of total arsenic removal from natural water sample. At 100 mg L−1 of NC-MA/L-MG 70% of arsenic was removed for 30 min and more than 99% after 60 min. Fast arsenic uptake was obtained with MC-O/L-MG adsorbent: 70% removal was achieved for 6 min and after 14 min the total arsenic concentration was below the WHO limit prescribed for arsenic in drinking water. This results showed that at even high concentration of interfering ion (Table 3) fast removal of As(V), present at low level, could be efficiently removed from natural water.

3.3.5

3.3.5 Regeneration versus safety

Experiments related to regeneration study with respect to both NC-MA/L-MG and MC-O/L-MG adsorbents, and related disposal issue of spent alkali desorption solution are given in Supplementary data.

3.3.6

3.3.6 Future challenges

The adsorption performances, comparable/higher than that of the most common referenced material, observed for well-dispersed/manipulative NC-MA/L-MG and MC-O/L-MG adsorbents result from the contribution of two factors: (i) design of surface modification with branched organic structure with number of amino terminal groups and (ii) development of controllable method for magnetite precipitation. The main goal of presented study was devoted to understanding of scientific principles related to influences of the parent media size, NC versus MC, on the adsorption properties of produced adsorbents. Also, investigation on the influences of chemical structure/properties of the modified precursor surface on the optimization of the method for magnetite precipitation was an additional goal of this study. Obviously, crystal structure of precipitated magnetite, exposed surface and crystal faces, had pronounced influence on adsorbent properties and strongly depends on experimental condition and NC/MC surface properties. It was shown that in a moderately acidic/neutral condition arsenic species are strongly bonded to iron oxide surface. EXAFS analysis suggested predominant formation of bidentate binuclear corner-sharing complexes (2C) for As(V), with dominant complexes formed between AsO43− tetrahedron and two FeO6 octahedra on {1 0 0} surfaces of magnetite (Liu et al., 2015). Also, calculation using density functional theory (DFT) predicted the relative energies and geometries of surface complexes using density functional theory showed higher stability of 2C complex for 55 kJ mol−1 in relation to 2E edge-sharing bidentate complex (Sherman and Randall, 2003). According to this, good adsorption performances of the novel adsorbents were related to proper NC and MC modification, and controllable MG precipitation which causes beneficial facet evolution at MG surface. As a result, high affinity and efficient arsenate removal from water, as well as acceptable kinetics with possibility for adsorbent use in the processes of natural water remediation even in the presence of competing ions were obtained.

Also, exceptional adsorption performances and magnetic properties of MG nanoparticles, precipitated on modified NC and MC surfaces, provide an opportunity for practical application and easy separation from aqueous solution by applying external magnetic field. Except this, results of reusability study showed low decrease in adsorption efficiency after five adsorption/desorption cycles (Supplementary data; Section 3.3.5). No measurable iron ion was detected in water during repeated cycles of adsorption. Therefore, MG loaded NC-MA/L and MC-O/L adsorbents could be effectively used for long time of exploitation. Using a proper selection of technology for treatment of spent waste alkali concentrated arsenic provides safe technology for arsenic removal (Supplementary data; Section 3.3.5). In that way, the three main goals were realized by understanding/explaining of scientific principles related to the influences of parent media size, NC versus MC, chemical structure/properties of organic functionalities bonded on NC and MC surface, and MG precipitation method on adsorption performances of novel synthesized adsorbents. Additionally, a number of idea/strategy were designed based on presented results/conclusions, and future work will be focused on development/production of woven nano/-micro cellulose MG modified hybrid textile materials.

4

4 Conclusion

Present study demonstrated that parent media size, functionalities/architecture of organic modifying structure and nanohybrid material have significant influence on adsorption properties of physically different materials. Compared to unsupported MG, two novel designed nano and micro cellulose supported adsorbents with organic support (amino functionalized) and inorganic nanohybrid precipitated adsorptive material (magnetite) exhibited favorable adsorption properties toward arsenate. Different geometry (aspect/ratio) and morphological/textural properties caused better adsorptive properties of NC-MA/L-MG, i.e. nearly five times higher adsorption capacity (85.3 vs 18.5 mg g−1) in comparison with MC-O/L-MG counterpart. Such results reflect the significance of increased number of amino surface bonding sites, present at NC-L surface, which are susceptible for effective MG precipitation. The results of kinetic study, for both adsorbents, were well fitted with the PSO and WM kinetic models. However, MC-O/L-MG demonstrated significantly faster kinetics, around three times higher value of K2, 0.541 vs 0.189 g mg−1 min−1, indicating better availability of surface active sites in MC-O/L-MG. Obtained rate constant in the presence of the MC-O/L-MG mainly exceeds the values found in the literature, and capacity of NC-MA/L-MG is similar/higher to literature data which confirms high applicability of both adsorbents. Iso-conversional method applied on kinetic data of both adsorbents showed that MC-O/L-MG was preferable material due to lower dependence of Ea on the extent of adsorption. Additionally, both adsorbents showed applicability in a competitive condition, reusability and possibility for separation of saturated adsorbent by magnetic field.

Acknowledgment

The authors acknowledge financial support from Ministry of Education, Science and Technological developments of the Republic of Serbia, Projects Nos. OI 172057 and III45019.

References

  1. , , . Removal of heavy metals from their aqueous solutions through adsorption onto natural polymers. Carbohydr. Polym.. 2011;84:454-458.
    [CrossRef] [Google Scholar]
  2. , , , , , . Phosphate removal and recovery from water using nanocomposite of immobilized magnetite nanoparticles on cationic polymer. Environ. Technol.. 2016;37:2099-2112.
    [CrossRef] [Google Scholar]
  3. , , . Removal of organic pollutants from water by modified cellulose fibres. Ind. Crops Prod.. 2009;30:93-104.
    [CrossRef] [Google Scholar]
  4. , , , . Modified cellulose fibres for adsorption of organic compound in aqueous solution. Sep. Purif. Technol.. 2006;52:332-342.
    [CrossRef] [Google Scholar]
  5. , , , . Removal of arsenic(V) from spent ion exchange brine using a new class of starch-bridged magnetite nanoparticles. Water Res.. 2011;45:1961-1972.
    [CrossRef] [Google Scholar]
  6. , , , . Synthesis and characterization of multi-carboxyl-functionalized nanocellulose/nanobentonite composite for the adsorption of uranium(VI) from aqueous solutions: kinetic and equilibrium profiles. Chem. Eng. J.. 2015;273:390-400.
    [CrossRef] [Google Scholar]
  7. , , . Cellulose-based anion exchanger with tertiary amine functionality for the extraction of arsenic(V) from aqueous media. J. Environ. Manage.. 2010;91:2201-2207.
    [CrossRef] [Google Scholar]
  8. , , , , . Development of an amino functionalized glycidylmethacrylate-grafted-titanium dioxide densified cellulose for the adsorptive removal of arsenic(V) from aqueous solutions. Chem. Eng. J.. 2012;209:362-371.
    [CrossRef] [Google Scholar]
  9. , . Assessing of lead(III) capturing from contaminated wastewater using ligand doped conjugate adsorbent. Chem. Eng. J.. 2016;289:65-73.
    [CrossRef] [Google Scholar]
  10. , . Solid phase sensitive palladium(II) ions detection and recovery using ligand based efficient conjugate nanomaterials. Chem. Eng. J.. 2016;300:264-272.
    [CrossRef] [Google Scholar]
  11. , . Ring size dependent crown ether based mesoporous adsorbent for high cesium adsorption from wastewater. Chem. Eng. J.. 2016;303:539-546.
    [CrossRef] [Google Scholar]
  12. , . A novel facial composite adsorbent for enhanced copper(II) detection and removal from wastewater. Chem. Eng. J.. 2015;266:368-375.
    [CrossRef] [Google Scholar]
  13. , , , . Removal of trace arsenic(V) and phosphate from water by a highly selective ligand exchange adsorbent. J. Environ. Sci.. 2011;23:1947-1954.
    [CrossRef] [Google Scholar]
  14. , , , , , , . Evaluating of arsenic(V) removal from water by weak-base anion exchange adsorbents. Environ. Sci. Pollut. Res.. 2013;20:421-430.
    [CrossRef] [Google Scholar]
  15. , , . Rapid column-mode removal of arsenate from water by crosslinked poly(allylamine) resin. Water Res.. 2009;43:1229-1236.
    [CrossRef] [Google Scholar]
  16. , , , , , . Encapsulation of cesium from contaminated water with highly selective facial organic–inorganic mesoporous hybrid adsorbent. Chem. Eng. J.. 2016;291:128-137.
    [CrossRef] [Google Scholar]
  17. , , , , , . Efficient arsenic(V) removal from water by ligand exchange fibrous adsorbent. Water Res.. 2012;46:5541-5550.
    [CrossRef] [Google Scholar]
  18. , , , , , . Arsenate removal from water by a weak-base anion exchange fibrous adsorbent. Water Res.. 2008;42:689-696.
    [CrossRef] [Google Scholar]
  19. , , , , , , . A reliable hybrid adsorbent for efficient radioactive cesium accumulation from contaminated wastewater. Sci. Rep.. 2016;6:19937.
    [CrossRef] [Google Scholar]
  20. , . Forward modeling of metal complexation by NOM: II. Prediction of binding site properties. Environ. Sci. Technol.. 2011;45:3202-3209.
    [CrossRef] [Google Scholar]
  21. , , , . Preparation and characterization of nano-magnetic cellulose with fast kinetic properties towards the adsorption of some metal ions. Chem. Eng. J.. 2012;191:22-30.
    [CrossRef] [Google Scholar]
  22. Dufresne, A., 2012. Nanocellulose: from nature to high performance tailored materials. De Gruyter Germany.
  23. , , , . The removal of heavy metal cations by natural zeolites. J. Colloid Interface Sci.. 2004;280:309-314.
    [CrossRef] [Google Scholar]
  24. , , . Insights into the modeling of adsorption isotherm systems. Chem. Eng. J.. 2010;156:2-10.
    [CrossRef] [Google Scholar]
  25. , . Vibrational studies of metal-ethylenediamine thiosulfates – I. Infrared and Raman spectra of the tris-ethylenediamine thiosulfates MII(en)3S2O3(MII = Zn, Cd, Fe, Ni Co, Mn) and some of their N-deuterated analogues. Spectrochica Acta. 1976;32:327-336.
    [CrossRef] [Google Scholar]
  26. , . Eradicating erroneous Arrhenius arithmetic. Thermochim. Acta. 2003;399:1-29.
    [CrossRef] [Google Scholar]
  27. , , , , , . Kinetic evaluation using a pressure-jump relaxation technique. Environ. Sci. Technol.. 1997;31:321-326.
    [CrossRef] [Google Scholar]
  28. , . The theory of rate processes (Glasstone, Samuel; Laidler, Keith J.; Eyring, Henry) J. Chem. Educ.. 1942;19:249.
    [CrossRef] [Google Scholar]
  29. , , , , , . Calcium hydroxyapatite microfibrillated cellulose composite as a potential adsorbent for the removal of Cr(VI) from aqueous solution. Chem. Eng. J.. 2016;283:445-452.
    [CrossRef] [Google Scholar]
  30. , , , , . Removal of arsenic(V) by magnetic nanoparticle activated microfibrillated cellulose. Chem. Eng. J.. 2015;260:886-894.
    [CrossRef] [Google Scholar]
  31. , , , , , , . Heavy metal removal from water/wastewater by nanosized metal oxides: a review. J. Hazard. Mater.. 2012;211–212:317-331.
    [CrossRef] [Google Scholar]
  32. , , , , , . A novel biodegradable β-cyclodextrin-based hydrogel for the removal of heavy metal ions. Carbohydr. Polym.. 2013;97:496-501.
    [CrossRef] [Google Scholar]
  33. , , , , . Separation and determination of arsenic species in water by selective exchange and hybrid resins. Anal. Chim. Acta. 2011;706:191-198.
    [CrossRef] [Google Scholar]
  34. , , , , . Surfactant transition metal chelates. Colloids Surfaces A Physicochem. Eng. Asp.. 2007;302:186-196.
    [CrossRef] [Google Scholar]
  35. , , , , , , , . Nanocelluloses: a new family of nature-based materials. Angew. Chemie Int. Ed.. 2011;50:5438-5466.
    [CrossRef] [Google Scholar]
  36. , , , , . Adsorption of arsenate on synthetic goethite from aqueous solutions. J. Hazard. Mater.. 2006;136:281-287.
    [CrossRef] [Google Scholar]
  37. , , , , . Microfibrillated cellulose – its barrier properties and applications in cellulosic materials: a review. Carbohydr. Polym.. 2012;90:735-764.
    [CrossRef] [Google Scholar]
  38. , , , , . In situ growing directional spindle TiO2 nanocrystals on cellulose fibers for enhanced Pb(2+) adsorption from water. J. Hazard. Mater.. 2015;289:140-148.
    [CrossRef] [Google Scholar]
  39. , , , . Removal of arsenic contaminants with magnetic γ-Fe2O3 nanoparticles. Chem. Eng. J.. 2012;211–212:46-52.
    [CrossRef] [Google Scholar]
  40. , , , , , , , . Mechanism of arsenic adsorption on magnetite nanoparticles from water: thermodynamic and spectroscopic studies. Environ. Sci. Technol.. 2015;49:7726-7734.
    [CrossRef] [Google Scholar]
  41. , , . Preparation and properties of cellulose nanocrystals: rods, spheres, and network. Carbohydr. Polym.. 2010;82:329-336.
    [CrossRef] [Google Scholar]
  42. , , , , , , , . Preparation and characterization of nano-cellulose with new shape from different precursor. Carbohydr. Polym.. 2013;98:562-567.
    [CrossRef] [Google Scholar]
  43. , , , . A novel cellulose–manganese oxide hybrid material by in situ soft chemical synthesis and its application for the removal of Pb(II) from water. J. Hazard. Mater.. 2010;181:986-995.
    [CrossRef] [Google Scholar]
  44. , , , , , , , . Ultrasonic assisted arsenate adsorption on solvothermally synthesized calcite modified by goethite, α-MnO2 and goethite/α-MnO2. Ultrason. Sonochem.. 2014;21:790-801.
    [CrossRef] [Google Scholar]
  45. , , , , , , , . Arsenate adsorption on waste eggshell modified by goethite, α-MnO2 and goethite/α-MnO2. Chem. Eng. J.. 2014;237:430-442.
    [CrossRef] [Google Scholar]
  46. , , , , , , . Influence of drying method on the material properties of nanocellulose I: thermostability and crystallinity. Cellulose. 2013;20:2379-2392.
    [CrossRef] [Google Scholar]
  47. , , . Bacteria immobilization on neem leaves/MnFe2O4 composite surface for removal of As(III) and As(V) from wastewater. Arab. J. Chem.. 2019;12(8):3263-3288.
    [Google Scholar]
  48. , , . Kinetics of thermal transformations of precipitated magnetite and goethite. J. Therm. Anal. Calorim.. 2001;65:497-503.
    [CrossRef] [Google Scholar]
  49. , , , . Arsenic Pollution: A Global Synthesis. United Kingdom: John Wiley & Sons; .
  50. , , . NIR FT Raman spectroscopy – a rapid analytical tool for detecting the transformation of cellulose polymorphs. Cellulose. 2001;8:49-57.
    [CrossRef] [Google Scholar]
  51. , , , . New method for determining the degree of cellulose I crystallinity by means of FT Raman spectroscopy. Cellulose. 2005;12:223-231.
    [CrossRef] [Google Scholar]
  52. Schwaminger, S.P., García, P.F., Merck, G.K., Bodensteiner, A., Heissler, S., Günther, S., Berensmeier, S., 2015. On the nature of interactions of amino acids with bare magnetite nanoparticles. doi:http://dx.doi.org/10.1021/acs.jpcc.5b07195.
  53. , , . Iron Oxides in the Laboratory (2nd ed.). Weinheim: Wiley-VCH; .
  54. , , . “Model-free” kinetic analysis ? Thermochim. Acta. 2002;390:217-225.
    [Google Scholar]
  55. , , . Raman spectroscopic study of magnetite (FeFe2O4): a new assignment for the vibrational spectrum. J. Solid State Chem.. 2003;174:424-430.
    [CrossRef] [Google Scholar]
  56. , , , , , . Development, characterization and evaluation of iron-coated honeycomb briquette cinders for the removal of As(V) from aqueous solutions. Arab. J. Chem.. 2014;7:27-36.
    [CrossRef] [Google Scholar]
  57. , , . Surface complexation of arsenic(V) to iron(III) (hydr)oxides: structural mechanism from ab initio molecular geometries and EXAFS spectroscopy. Geochim. Cosmochim. Acta. 2003;67:4223-4230.
    [CrossRef] [Google Scholar]
  58. , . Unexplored possibilities of all-polysaccharide composites. Carbohydr. Polym.. 2013;95:697-715.
    [CrossRef] [Google Scholar]
  59. , , , . Functionalized nanocrystalline cellulose: smart biosorbent for decontamination of arsenic. Int. J. Miner. Process.. 2015;139:51-63.
    [CrossRef] [Google Scholar]
  60. , . Infrared and Raman Characteristic Group Frequencies: Tables and Charts. John Wiley & Sons; .
  61. , , , , , , , . Amino-functionalized magnetic cellulose nanocomposite as adsorbent for removal of Cr(VI): synthesis and adsorption studies. Chem. Eng. J.. 2014;241:175-183.
    [CrossRef] [Google Scholar]
  62. , , , . Sensing the structural differences in cellulose from apple and bacterial cell wall materials by Raman and FT-IR spectroscopy. Sensors. 2011;11:5543-5560.
    [CrossRef] [Google Scholar]
  63. , , , , , , , . Efficient arsenic removal by cross-linked macroporous polymer impregnated with hydrous iron oxide: material performance. Chem. Eng. J.. 2015;279:66-78.
    [CrossRef] [Google Scholar]
  64. , , , , , , , , , . Modified native cellulose fibers – a novel efficient adsorbent for both fluoride and arsenic. J. Hazard. Mater.. 2011;185:93-100.
    [CrossRef] [Google Scholar]
  65. , , , , , , , . Removal of lead from water by amino modified multi-walled carbon nanotubes. Chem. Eng. J.. 2011;173:855-865.
    [CrossRef] [Google Scholar]
  66. , , . Band assignments in the Raman spectra of celluloses. Carbohydr. Res.. 1987;160:113-129.
    [CrossRef] [Google Scholar]
  67. , , , . Fast and selective adsorption of As(V) on prepared modified cellulose containing Cu(II) moieties. Arab. J. Chem.. 2016;9(5):607-615.
    [Google Scholar]
  68. , , , , , , , . Synthesis and characterization of multi-amino-functionalized cellulose for arsenic adsorption. Carbohydr. Polym.. 2013;92:380-387.
    [CrossRef] [Google Scholar]
  69. , , , , , , , , . Adsorption of Cd(II), Cu(II) and Ni(II) ions by cross-linking chitosan/rectorite nano-hybrid composite microspheres. Carbohydr. Polym.. 2015;130:333-343.
    [CrossRef] [Google Scholar]
  70. , , , . FT Raman spectroscopic investigation of cellulose acetate. Cellulose. 2011;18:995-1003.
    [CrossRef] [Google Scholar]
  71. , , , . Preparation and evaluation of a magnetite-doped activated carbon fiber for enhanced arsenic removal. Carbon N. Y.. 2010;48:60-67.
    [CrossRef] [Google Scholar]
  72. , , , , . Synthesis of the cotton cellulose based Fe(III)-loaded adsorbent for arsenic(V) removal from drinking water. Desalination. 2009;249:1006-1011.
    [CrossRef] [Google Scholar]

Appendix A

Supplementary material

Supplementary data associated with this article can be found, in the online version, at http://dx.doi.org/10.1016/j.arabjc.2016.08.006.

Appendix A

Supplementary material

Supplementary data 1

Supplementary data 1

Show Sections